316
vii Reservoir Quality Prediction in Sandstones and Carbonates: An Overview Julie A. Kupecz Intevep, S.A. Los Teques, Venezuela and Kupecz and Associates, Ltd. Denver, Colorado, U.S.A. Jon Gluyas Monument Oil and Gas London, United Kingdom Salman Bloch Texaco E&P Technology Department Houston, Texas, U.S.A INTRODUCTION The accurate prediction of reservoir quality is, and will continue to be, a key challenge for hydrocarbon exploration and development. Prediction is a logical and critically important extension of the description and interpretation of geological processes. However, in spite of the profusion of publications on sandstone and carbonate diagenesis, relatively few articles illustrate the application of such studies to reservoir quality pre- diction. This Memoir represents the first attempt to compile worldwide case studies covering some predic- tive aspects of both siliciclastic and carbonate reservoir characteristics. We have attempted here to focus on the variability due to diagenetic effects in sandstones and carbonates, rather than on sedimentological effects, i.e., the presence or absence of a given reservoir. The chap- ters cover the spectrum of stages in the exploration- exploitation cycle (Table 1). The importance of reservoir quality in pay evalua- tion has been illustrated by Rose (1987), who analyzed an unnamed company’s exploration results over a 1-year period. Of 87 wildcat wells drilled, 27 were dis- coveries (31% success rate); incorrect predictions of the presence of adequate reservoir rocks were made in 40% of the dry holes. Importantly, the geologists believed that reservoir quality was the primary uncer- tainty in 79% of the unsuccessful wells. Similarly, a comparison of predrill predictions with postdrill results by Shell (Sluijk and Parker, 1984) indicated that reservoir quality was seriously overestimated, whereas hydrocarbon charge and retention predic- tions were more accurate. Although these statistics do not clearly separate drilling failure due to lack of potential reservoir from the lack of adequate reser- voir quality, it seems that although explorers are aware of the significance of reservoir quality predic- tion, generation of predictive models continues to be a formidable task. Accurate prediction of reservoir quality is needed throughout the entire “life cycle” of a reservoir (Snei- der, 1990). Proper assessment of reservoir quality must be continually refined, from prior to exploratory drilling, to discovery, during appraisal and develop- ment drilling, and throughout reservoir management. At the Exploration Stage, the main challenge is to assess and predict the reservoir facies, its geometry, and its distribution; reservoir porosity and permeability for use in petroleum reserves calculations; seismic charac- teristics; and migration pathways. In this Memoir, papers by Brown, Ehrlich et al., Evans et al., Gluyas, Gluyas and Cade, Gluyas and Witton, Primmer et al., Ramm et al., Sombra and Chang, Tobin, and Zem- polich and Hardie address various aspects of the assessment process. At the Appraisal, Planning, and Development Stages, it is necessary to understand and predict reservoir porosity, permeability, and reservoir distribution to

Reservoir Quality Prediction in Sand and Carbonates

Embed Size (px)

Citation preview

Page 1: Reservoir Quality Prediction in Sand and Carbonates

vii

Reservoir Quality Prediction in Sandstones and Carbonates:

An OverviewJulie A. Kupecz

Intevep, S.A.Los Teques, Venezuela

andKupecz and Associates, Ltd.Denver, Colorado, U.S.A.

Jon GluyasMonument Oil and Gas

London, United Kingdom

Salman BlochTexaco E&P Technology Department

Houston, Texas, U.S.A

INTRODUCTION

The accurate prediction of reservoir quality is, andwill continue to be, a key challenge for hydrocarbonexploration and development. Prediction is a logicaland critically important extension of the descriptionand interpretation of geological processes. However, inspite of the profusion of publications on sandstone andcarbonate diagenesis, relatively few articles illustratethe application of such studies to reservoir quality pre-diction. This Memoir represents the first attempt tocompile worldwide case studies covering some predic-tive aspects of both siliciclastic and carbonate reservoircharacteristics. We have attempted here to focus on thevariability due to diagenetic effects in sandstones andcarbonates, rather than on sedimentological effects, i.e.,the presence or absence of a given reservoir. The chap-ters cover the spectrum of stages in the exploration-exploitation cycle (Table 1).

The importance of reservoir quality in pay evalua-tion has been illustrated by Rose (1987), who analyzedan unnamed company’s exploration results over a 1-year period. Of 87 wildcat wells drilled, 27 were dis-coveries (31% success rate); incorrect predictions of thepresence of adequate reservoir rocks were made in40% of the dry holes. Importantly, the geologistsbelieved that reservoir quality was the primary uncer-tainty in 79% of the unsuccessful wells. Similarly, acomparison of predrill predictions with postdrill

results by Shell (Sluijk and Parker, 1984) indicatedthat reservoir quality was seriously overestimated,whereas hydrocarbon charge and retention predic-tions were more accurate. Although these statisticsdo not clearly separate drilling failure due to lack ofpotential reservoir from the lack of adequate reser-voir quality, it seems that although explorers areaware of the significance of reservoir quality predic-tion, generation of predictive models continues to bea formidable task.

Accurate prediction of reservoir quality is neededthroughout the entire “life cycle” of a reservoir (Snei-der, 1990). Proper assessment of reservoir quality mustbe continually refined, from prior to exploratorydrilling, to discovery, during appraisal and develop-ment drilling, and throughout reservoir management.At the Exploration Stage, the main challenge is to assessand predict the reservoir facies, its geometry, and itsdistribution; reservoir porosity and permeability foruse in petroleum reserves calculations; seismic charac-teristics; and migration pathways. In this Memoir,papers by Brown, Ehrlich et al., Evans et al., Gluyas,Gluyas and Cade, Gluyas and Witton, Primmer et al.,Ramm et al., Sombra and Chang, Tobin, and Zem-polich and Hardie address various aspects of theassessment process.

At the Appraisal, Planning, and Development Stages, itis necessary to understand and predict reservoirporosity, permeability, and reservoir distribution to

Page 2: Reservoir Quality Prediction in Sand and Carbonates

viiiTable 1. Overview of Chapters in This Memoir.

Stage inExploration-Exploitation Data/ Summary of

Author Cycle Location/Basin Reservoir Age Lithology Methodology ChapterBrown Exploration North Dakota, Mississippian Carbonates Wireline logs, Determination of influence of

Williston Basin, (limestone, cuttings descriptions, carbonate mineralogy, shale content, U.S.A. dolomite, temperature, numerical and fabric on loss of porosity with

argillaceous regression burial.carbonate)

Cabrera-G, Development Western Canada Devonian Carbonate Seismic Porosity prediction from Arestad, Sedimentary Basin (dolomite), multicomponent seismic data via Dagdelen, evaporite, geostatistical methods.and Davis shale

Cavallo Development West Virginia, Mississippian Carbonate Formation FMS logs with sidewall core, inte-and U.S.A.; Appalachian Microscanner (FMS) grated into depositional model Smosna Basin logs for ooid shoals. Used to predict opti-

mal location for development wells.

Erlich, Exploration; Examples from Miocene; Permian– Sandstone Petrographic Integration of PIA and porosity to Bowers, Development Thailand (Pattani Late Carboniferous, Image Analysis understand variations in permeability.Riggert, and Basin), Oklahoma, respectively. (PIA), mercury Prince U.S.A. (Cherokee Basin) porosimetry

Evans, Variable; N/A N/A Sandstone Modeling of Modeling effects of geological Cade, and Overview of empirical data processes that affect permeability Bryant permeability (porosity, lithology) (burial, cementation) to calculate

prediction changes in permeability.

Gluyas Exploration Norwegian Central Late Jurassic Sandstones Petrography; porosity, Risking of porosity evolution models Graben permeability data for predrill porosity prediction.

Gluyas Exploration Worldwide Permian to Sandstones Integration of Porosity–depth relationship for and published data Pleistocene (quartz, experimental, prediction in uncemented sandstones Cade feldspar) petrographic, and gives maximum porosity baseline to

porosity data compare cement volumes and (worldwide) cemented ss porosity.

Gluyas Exploration Southern Red Sea, Miocene Sandstone Petrography, Case study of predrill reservoir and offshore Yemen burial and thermal quality prediction.Witton history, provenance,

depositionalenvironment

Love, Development N. Germany; Permian Carbonate Statistics; neural Statistical relationships of Strohmenger, Southern (dolomite; networks; core; geological data for prediction of Woronow, and Zechstein calcitized well logs; predrill reservoir quality.Rochenbauch Basin dolomite) structural data;

geochemistry

Page 3: Reservoir Quality Prediction in Sand and Carbonates

ix

Major Development; Permian Basin, Permian Carbonate Petrography, Determination of “flow units” and Reservoir west Texas and SE (dolomite) well logs, capillary controlled by depositional facies Holtz Management New Mexico, U.S.A. pressure data and diagenetic alteration;

cores; well-logs; quantification of bypassed oil in porosity and low-permeability flow units and permeability data; heterogeneous flow units.production history

Mountjoy Development Western Canada Devonian Carbonate Petrography Controls of depositional facies and and Sedimentary Basin (dolomite, diagenesis on pore systems and Marquez limestone) reservoir continuity; effects of

dolomitization on pore types and reservoir character; comparison of reservoir characteristics of limestone vs. dolomite at depth.

Primmer, Exploration Worldwide data Variable; Sandstones Depositional Subdivision into five “styles” of Cade, Evans, base predominantly environment, diagenesis via relationship between Gluyas, Mesozoic and composition, detritial composition, burial depth, Hopkins, younger maximum burial temperature, cement type.Oxtoby, time; fluid inclusions, Smalley, stable isotopes, andWarren, and organic maturation Worden where available

Ramm Exploration Norwegian Central Late Jurassic Sandstones Petrography; Porosity prediction by prediction of Graben fluid inclusions composition, texture, and

microquartz coatings that inhibit quartz cementation.

Smosna Exploration Pennsylvania, Devonian Sandstones Petrography Prediction of reservoir potential of and U.S.A.; Appalachian (litharenites and range of depositional facies.Bruner Basin sublitharenites)

Sombra Exploration Brazil: Santos, Late Jurassic– Sandstones Petrography; Time Depth Index (TDI) to quantify and Campos, Espiritu Tertiary porosity vs. depth influence of burial history onChang Santo, Cumuruxatiba, porosity evolution.

Reconcavo, Sergipe, Alagoas, and Potiguar basins

Tobin Exploration Examples from Triassic; Paleocene– Sandstones, Outcrop Decision Tree to classify outcrop for China, Myanmar, Eocene; Jurassic, carbonates, risk assessment.Turkey respectively respectively

Zempolich Exploration Venetian Alps, Middle Jurassic Carbonate Outcrop; Field mapping of dolomite distributionand Italy (limestone, petrography; for information on size andHardie dolomite) geochemistry distribution of dolomite bodies and

evidence for fluid pathways. Studyof progressive textural modificationfor prediction of reservoir-gradeporosity, permeability.

Page 4: Reservoir Quality Prediction in Sand and Carbonates

x

determine the location and optimal number of devel-opment wells, as well as to estimate economic produc-tion cutoff values, hydrocarbon pore volumes,recoverable reserves, and production rates (Sneider,1990). By understanding controls on the degree ofreservoir heterogeneity and distribution of flow units,a more accurate understanding and predictability ofinterwell connectivity and fluid-flow pathways can begained (Tyler et al., 1984; Ebanks, 1990; Kerans et al,1994; Stoudt and Harris, 1995; Tinker, 1996). Studies atthe development scale in this Memoir are provided byLove et al., Smosna and Bruner, and Cavallo andSmosna. Prediction of permeability is addressed byEvans et al., Gluyas and Witton, and Erlich at al. Theevolution of permeability during diagenesis isaddressed in this Memoir by Zempolich and Hardie,and Mountjoy and Marquez.

At the Reservoir Management Stage, predictability ofdiagenetic patterns that control reservoir quality isused to identify bypassed and uncontacted pay, and intertiary recovery planning and modification. Identifi-cation of bypassed oil and quantification of remaininghydrocarbons is addressed in this volume by Majorand Holz.

COMPARISON OF SANDSTONES ANDCARBONATES: REASONS FOR

SIMILARITIES AND DIFFERENCES INPREDICTIVE APPROACHES

There are some similarities and many differencesbetween siliciclastics and carbonates, both in theirdepositional characteristics and in the way in whichthey respond to physical and chemical conditions dur-ing burial and lithification. Clearly, the total of thedepositional and diagenetic effects control the final“reservoir-quality” product. In the following discus-sion, we compare both similarities and differencesbetween sandstones and carbonate rocks under theguise of three headings: depositional controls, diage-netic controls, and resultant pore types.

Depositional Processes and Controls on Reservoir Quality Prediction

In contrast to siliciclastics, the generation and depo-sition of most carbonates is controlled by biologicalactivity (~90%; Moore, 1989); sand generation anddeposition is much less influenced by life. The signifi-cance of biological control on carbonate accumulationis that thickness and depositional properties of carbon-ates can form independently of allochthonous sedi-ment supply. Certain prerequisites must be met forcarbonates to form (e.g., temperature, light, salinity,and the availability of nutrients), which will controltheir geographical location as well as their environ-ments of deposition. As a result, most carbonates arelimited to shallow, tropical marine depositional set-tings. Adding complexity to reservoir quality predic-tion is that carbonate-producing organisms haveevolved through time (e.g., Wilson, 1975; James, 1978).

In contrast, sand is derived mainly from erosion of aparent source and is transported to its site of deposi-tion by physical processes. Physical parameters ofsandstones (grain size, sorting, roundness, etc.) areused to understand and predict depositional processesand environments in which they were deposited.Some carbonate depositional environments are alsostrongly influenced by hydrologic controls, and result-ing facies will have similar depositional characteristicsto siliciclastic sandstones (e.g., bars, shoals, beaches,dunes, tidal flats, tidal channels, tidal deltas, and basin-margin sediment gravity flow deposits; Scholle et al.,1983, and references therein).

The similarities and differences between carbonateand siliciclastic sedimentology are reflected in similar,yet contrasting, concepts of sequence stratigraphy.The concepts of carbonate sequence stratigraphy aresummarized by Sarg (1988), Schlager (1992), andHandford and Loucks (1993) and can be compared tosandstone sequence stratigraphy (e.g., Mitchum, 1977;Mitchum et al., 1977; Vail et al., 1977; Posamentier etal., 1988; Van Wagoner et al., 1988, 1990; among oth-ers). Large-scale stratal geometries of siliciclastic sedi-ments (onlap, downlap, toplap, etc.) are also thefundamental geometries of carbonate depositionalsequences. The relative volumetric importance of dif-ferent systems tracts, however, is different for sandsvs. carbonates.

Siliciclastics are controlled by physical sedimentsupply. During relative highstand of sea level, mostcoarse-grained clastics are “trapped” in fluvial sys-tems and are not deposited in marine settings. Duringrelative lowstands of sea level, coarse-grained sedi-ments are able to bypass the shelf to be deposited inbasinal marine settings. Therefore, lowstand systemstracts (LST) generally contain the most volumetricallyabundant deposits of coarse-grained siliciclastics inpetroleum basins. In contrast, the most significant fac-tor for carbonate deposition is the inundation of shal-low carbonate platforms (Sarg, 1988; Schlager, 1992;Handford and Loucks, 1993). As a result, during rela-tive highstands of sea level, carbonates will be able togenerate and accumulate the most significant quanti-ties of sediment, varying according to relative rates ofsediment production, accumulation, and sea level rise(Sarg, 1988). Therefore, highstand systems tract (HST)deposits are generally the most volumetrically signifi-cant for carbonates. During relative sea level lowstands, carbonate deposition is generally geographically and volumetrically restricted and lesssignificant, although allochthonous slope-derivedmaterial and autochthonous deposits may be locallyimportant.

The fundamental differences between the way inwhich carbonates and siliciclastics accumulate and areeroded and redeposited during a highstand–lowstandcycle have a major effect on the evolution of reservoirquality. Typically, sands deposited during highstandswill suffer erosion and redeposition down systemstract as sea level falls, but the modification of the sedi-ment is dominantly physical rather than chemical.Highstand carbonate deposits are unlikely to suffer

Page 5: Reservoir Quality Prediction in Sand and Carbonates

the same fate. Exposure during sea level fall will bedominated by dissolution and reprecipitation ratherthan physical reworking of sediment. Depending onthe climate, time, and magnitude of exposure, karstifi-cation, dolomitization, and evaporite precipitation canoccur, all of which will result in a profound modifica-tion of reservoir quality.

In summary, differences in depositional controls,depositional and sequence stratigraphic settings, andsequence stratigraphic concepts between sandstonesand carbonates necessitate that approaches to facies-and reservoir-quality prediction in sandstones vs. carbonates, although fundamentally similar, must alsobe specific and characteristically different.

Mineralogy, Diagenesis, and Reservoir-QualityModification

Mineralogy

The second fundamental difference between car-bonates and sandstones is mineralogy and the way inwhich the mineralogy both responds to and, indeed,controls diagenesis. Mineralogy of sandstones,although variable, commonly consists of grains thatare chemically stable in the near-surface depositionalenvironment. Although dissolution of feldspars andlithic fragments can be locally important (Heald andLarese, 1973; Milliken et al., 1989; Milliken, 1992; Blochand Franks, 1993; among others), changes in porosityand permeability are not generally sufficient to signifi-cantly improve the overall quality of a reservoir(Bloch, 1994).

Carbonate sediments, in contrast, are composed of asmall variety of minerals that are highly susceptible tochemical alteration, recrystallization, and dissolution(e.g., aragonite, Mg-calcite, calcite, and dolomite ofvarying stoichiometry). The effects of carbonate min-eral instability on reservoir quality may be accentu-ated by the tendency of highstand carbonate systemsto be exposed during falling sea level. The water:rockratio during meteoric flushing and repeated seawaterinundation is clearly much larger than that likely to beexperienced during burial conditions. Consequently,there is significant potential for diagenetic modifica-tion before and throughout burial, often with multiplediagenetic events superimposed, and a continual mod-ification of reservoir quality.

Meteoric Diagenesis

Subaerial exposure, meteoric diagenesis, and subse-quent porosity evolution in carbonates have beenaddressed by Saller et al. (1994) and Budd et al. (1995).Among the most significant factors that determine themagnitude of carbonate porosity redistribution are thefollowing: mineralogy, existing pore networks, depo-sitional facies and stratigraphy, climate, the reactivepotential of the groundwater, duration of exposure,hydrologic systems, size and topography of theexposed area, magnitude of base-level change, andtectonic setting. Exposure of carbonates can be mani-fest in two important diagenetic processes, karstifica-tion and meteoric cementation, with significant

redistribution of porosity and permeability takingplace from the time of exposure throughout burial.

Studies of modern and ancient carbonate rocks sub-jected to exposure and meteoric diagenesis have docu-mented the variability of the cementation process and itsvariable effectiveness. Enos and Sawatsky (1981) docu-mented the high but variable nature of initial porosity ofmodern carbonate sediments (values ranging from 40%to 78%), and inferred that early diagenetic processesare responsible for the significant loss of preburialporosity (~20% loss in porosity) in analogous facies ofnearby Pleistocene rocks. Budd et al. (1993) estimatedthat precompaction meteoric cements account for 3–37vol. % in grainstones. However, Halley and Beach(1979) and Scholle and Halley (1985), based on studiesof Holocene and Pleistocene sediments of Florida andthe Bahamas, have claimed that porosity loss is slightduring mineralogical stabilization, and that secondaryporosity developed during early cementation pre-serves the overall magnitude of preburial porosity.These examples highlight the problem of uncertaintyin preburial porosity prediction in carbonates.

Meteoric diagenesis in sandstones is a controversialtopic. Much of the controversy has focused on the gen-eration of secondary porosity. The complexity of theprocesses involved precludes any a priori assumptionsas to the quantitative importance, or even presence, ofsecondary and enhanced porosity associated withmeteoric diagenesis (Bloch, 1994). Furthermore, identi-fication and quantification of secondary porosity oftenrely on subjective criteria. Even when positive evi-dence exists, such as partially dissolved grains and/orcements, it may be difficult to prove a meteoric originfor mineral dissolution. Giles and Marshall (1986), in areview of secondary porosity in sandstones, made aplausible case for the involvement of meteoric waterdissolution in some settings. More recently, Emery etal. (1990) have furnished strong evidence using a com-bination of wireline log, core analysis, thin section, iso-tope geochemical, and seismic acoustic impedancedata to highlight meteoric water dissolution of sand-stones beneath an unconformity. The possibility thatmeteoric water can penetrate deep into a basin andstill influence the course of diagenesis has beendemonstrated from analysis of the oxygen and hydro-gen/deuterium isotope ratios in authigenic minerals(Gluyas et al., 1997).

Marine Diagenesis

Active marine cementation, the occlusion of porosity,and the modification of pore types in various moderncarbonate marine depositional settings have been docu-mented by many workers (Bathurst, 1975, and refer-ences therein). Attesting to its economic importance,the significance of marine cementation in ancient car-bonate reefs and buildups has been documented in avast number of studies (e.g., Playford, 1980; and inbooks edited by Bebout and Loucks, 1977; Toomey,1981; Schneidermann and Harris, 1985; Schroeder andPurser, 1986; and Monty et al., 1995; among others).The variability and magnitude of marine diageneticeffects on reservoir quality in carbonates are illustrated

xi

Page 6: Reservoir Quality Prediction in Sand and Carbonates

xii

by Walls and Burrowes (1985), who documented that15% to 70% of total porosity in Devonian reefs ofCanada has been occluded by marine cement. Keranset al. (1986) estimated that in Devonian reefs of theCanning Basin, Australia, radiaxial and microcrys-talline marine cements each locally comprise 20–50%of the reef by volume.

There is no well-defined division of sandstone dia-genesis into marine vs. nonmarine. Admittedly, mete-oric water-influenced mineral dissolution has beenmuch investigated because of the potential effect onreservoir quality improvement, as discussed above.However, near-surface precipitation processes canoccur in a variety of environments (fluvial, marine,evaporitic, etc.). Carbonates, sulfates, and possiblyhalite tend to be the most important. These cements,however, rarely completely destroy the pore system ina large sand body. Moreover, because it is common forsuch cements (particularly carbonate) to form concre-tions, layers, or irregular masses, the effect on reser-voir quality is often best represented as a reduction inthe net (petroleum) pay thickness of a reservoir ratherthan the average effect on porosity (Bjørkum andWalderhaug, 1990). The diagenetic processes control-ling these near-surface reactions are relatively wellunderstood, and commonly involve bacterial destruc-tion of organic matter in oxic, suboxic, and anoxic porewaters (Berner, 1980). However, although the processis well understood, methods are as yet unavailable forpredicting the volume of syndepositional/early diage-netic cements in sandstones awaiting the drill bit.

Burial Diagenesis

Numerous diagenetic studies have documentedthat abundant cementation of carbonates occurs in theburial realm, which reduces or occludes any remain-ing porosity. The use of cathodoluminescence stratig-raphy (e.g., Meyers, 1991; among others) has beenshown to be an extremely useful tool for identifyingand correlating generations of cement. Cathodolumi-nescence techniques have allowed workers to correlatephases of cementation to geochemical environments(e.g., meteoric, marine, burial) and then to estimatevolume of cement precipitated during the various dia-genetic phases. Grover and Read (1983) concludedthat major, but variable, cementation has occurredunder burial conditions in the Middle Ordovician ofVirginia (U.S.A.), with 3–45 vol. % of cement duringshallow burial (≤3 km) and 50–95% during deep bur-ial. Meyers and Lohmann (1985), in their study of theMississippian limestones of New Mexico (U.S.A.), esti-mated that approximately 60% of total cement wasrelated to shallow-burial, marine phreatic processes,while approximately 40% was related to burial deeperthan 1 km. Dorobek (1987) estimated that approxi-mately 32% of the total cement in the Silurian–Devon-ian Helderberg Group of the central Appalachians(U.S.A.), was precipitated during shallow burial, withcementation by deep burial fluids occluding allremaining porosity. Using chemical, isotopic, and pet-rographic analysis, Prezbindowski (1985) estimatedthat 14 vol. % cement in the Cretaceous Stuart City

reefs of Texas (U.S.A.) was due to marine cementation,7 vol. % to near-surface, meteoric cementation, and 9vol. % as the result of burial cementation.

Burial diagenesis and its effects on the quality ofpetroleum reservoirs is a much-researched topic. Therange of minerals that can reduce the quality of a reser-voir is large: quartz, carbonate minerals, clays, zeolites,and others (Primmer et al., this volume). The applica-tion of quantitative petrographic, geochemical, and iso-topic analyses to authigenic minerals during the pastdecade has allowed scientists to date minerals, deter-mine the temperature of precipitation, and characterizethe pore waters from which precipitation occurred(e.g., Emery and Robinson, 1993; Williams et al., 1997).When such data are coupled with analyses of thermaland burial history information, powerful descriptionsof diagenetic process have emerged (Glasmann et al.,1989; Kupecz and Land, 1991; Robinson and Gluyas,1992; Hogg et al., 1993; Walderhaug, 1994). However,some key questions remain unanswered (e.g., thereappears to be too little connate water in sediments toredistribute the observed cement volumes in the timeavailable to the process). Essentially, there is insuffi-cient knowledge at present to determine the controls(source/transport/precipitation of solutes) on the dia-genetic evolution of sandstones. As for transport itself,there are advocates of lateral fluid flow, advection, anddiffusion as the major harbingers of cementing fluids.This paucity of quantitative knowledge means thatprocess-based predictive methodologies are few, andempiricism remains the prime tool for prediction ofreservoir quality.

Dolomitization

Dolomitization can occur during essentially synsed-imentary replacement or cementation of precursor car-bonate and can continue throughout the burial realm.A spectrum of environments have been proposed bymany (summarized by Land, 1980, 1982, 1985, 1986;Morrow, 1982, among others). Work in recent yearshas highlighted the fact that nonstoichiometricdolomites are susceptible to recrystallization (e.g.,Kupecz et al., 1993), and that recrystallization is com-monly associated with a progressive increase in crystalsize (Kupecz and Land, 1994). The significance ofdolomitization for reservoir quality is that an increasein crystal size (either during dolomitization of amicrite-dominated precursor or during dolomiterecrystallization) and/or the rearrangement of touch-ing pore space is generally associated with increasedpermeability (Lucia et al., 1995; Zempolich andHardie, this volume). Because of the complexity of thedolomitization process and the potential for continueddolomite modification, prediction of reservoir qualitywill have inherent uncertainties.

Variability in Pore Types and Reservoir Quality Prediction

Pore types and their distribution are fundamentallydifferent in sandstones and carbonates (e.g., Choquetteand Pray, 1970, their table 1). The dominant primary

Page 7: Reservoir Quality Prediction in Sand and Carbonates

pore type in sandstones is interparticle, regardless ofdepositional setting, with the pore diameter and pore-throat size a function of grain size and sorting (e.g.,Evans et al., this volume). Cementation by quartz (asolid grain coating) and mechanical compaction willreduce pore and pore-throat dimensions, but the poretypes remain essentially the same. The process of com-paction or quartz cementation can proceed to lowporosity levels without altering the relationshipbetween porosity and permeability. Only when cemen-tation proceeds to the point where pore coordinationnumber declines (i.e., pore throats are being closed off)is there a major change in the poroperm relationship,with permeability falling to very low levels. Typicallyfor a clean quartzose, medium-grained sandstone,porosity can be reduced to ~10% before the poropermrelationship declines. The porosity threshold will behigher for finer grained and more poorly sorted sands.Disruption of the pore network can occur at muchhigher porosity levels, where a mineral plugs poresrandomly or creates “furry” microporous grain coats.Typically, carbonate minerals or clusters of kaoliniteplatelets plug pores, while chlorite and illite are com-mon as clay coats with much trapped microporosity.Grain dissolution may result in moldic and micro-moldic porosity.

Carbonate primary pore types are highly variable,with their shapes and sizes having little relation toenergy, grain size, or sorting. Diagenetic modifica-tion of carbonate pore types adds additional com-plexity, with the resulting “ultimate” pore typevarying widely (Choquette and Pray, 1970). Pores incarbonate rocks can range in size from <1 µ to cav-erns >100 m in diameter, and may be juxtaposedwithin the same rock unit. The complexity of poros-ity in carbonates is the result of many factors, whichinclude the variable dimensions of sedimentary car-bonate particles, the variability of skeletal pores,partial to total occlusion of pores by internal sedi-ment or cement, creation of secondary pores [fabricselective or fabric independent, and of highly vari-able dimensions (e.g., breccias)], dolomitization, andrecrystallization (e.g., Murray, 1960; Choquette andPray, 1970). Because of the combination of biologicaland physical depositional processes, and diageneticoverprint of metastable chemical deposits, buriedcarbonates tend to have a greater heterogeneity ofporosity and permeability than do buried sand-stones and, as a result, generally have a greateruncertainty in prediction of average porosity.

PRESENT AND EMERGINGMETHODOLOGIES OF RESERVOIR

QUALITY PREDICTION

Current geological approaches to predict porosityand permeability in reservoirs prior to drilling rangebetween theoretical chemical models and purelyempirical models (Byrnes, 1994). Regardless of theapproach, to be useful from a practical point of view, a

predictive technique must meet a number of criteria(Bloch and Helmold, 1995):

1. Sufficient accuracy must be achieved from a lim-ited number of input parameters that can be esti-mated prior to drilling;

2. Prediction must be possible for a wide range oflithologies occurring in different geologic settings;

3. Permeability should be predicted independentlyof porosity to reduce the margin of error;

4. Although current understanding of processesresponsible for porosity preservation, destruc-tion, and enhancement is limited, the predictivemodel should at least implicitly account for themost important processes that take place duringsediment burial;

5. For production and exploration purposes, theapproach should be applicable on the reservoirscale, field scale, and subbasin scale. Basin-scalepredictions are adequate for basin modeling, butnot for the drilling of specific targets; and

6. The technique should be flexible, so that when itis not adequate by itself, reasonable accuracy canstill be achieved by using it with anotherapproach.

Choice of approach depends upon the type of antic-ipated reservoir rock and the amount of informationavailable. In mature areas where cores and logs pro-vide a calibration data set, the empirical approachesmay prove best. This is especially true with fielddevelopment prediction. In undrilled basins or targets,some aspect of theoretical relationships must be used,because there are no empirical data. In some cases, theuncertainty of the prediction will be large. This uncer-tainty should be related along with the predictivevalue so the value of the prediction can be correctlyassessed.

Sandstones

Process-Oriented Models

Process-oriented models (or chemical reaction pathmodels) do not meet some of the above criteria (mostnotably the first criterion). Such models are useful insimulating formation of some cements and diageneticsequences in simple compositional systems (Bruton,1985; Harrison, 1989; Harrison and Tempel, 1993), butare not yet capable of quantifying changes in porosityand permeability (Surdam and Crossey, 1987;Schmoker and Gautier, 1988; Meshri, 1989; Harrisonand Tempel, 1993). The limitations of these modelsinclude the following: (1) uncertainties in thermody-namic and kinetic data used in the reaction path calcu-lations (Surdam and Crossey, 1987; Meshri, 1989;Harrison and Tempel, 1993), (2) inaccuracies in paleo-hydrologic reconstructions, (3) inability to quantifymass transfer processes and the effect of theseprocesses on reservoir quality (Harrison and Tempel,1993), and (4) lack of feedback between compactionalporosity loss and mineral reactions (Harrison and

xiii

Page 8: Reservoir Quality Prediction in Sand and Carbonates

xiv

Tempel, 1993). Despite their limitations, chemical reac-tion path models are useful, as they attempt to explainmechanistically what is occurring during porosityevolution and, thus, are helpful in identifying criticalissues for further scientific studies of porosity evolu-tion (Waples and Kamata, 1993).

Empirical Models

By contrast, empirical techniques can be a powerfulpredictive tool, but their effectiveness is to a largeextent a function of availability and quality of calibra-tion data sets. Reservoir quality prediction is no excep-tion to the general rule that the fewer the calibrationdata, the less certain the prediction. The statement ofWaples et al. (1992, p. 47), that maturity models “aresimply too weak at present to allow us to carry outhighly meaningful modeling unless our input is con-strained by measured data” is also true of predictingreservoir quality.

In frontier areas, where data are sparse or not avail-able, only comparative analogs can be used. If surfaceoutcrops are available, the approach proposed by Tobin(this volume) can significantly assist in assessing poten-tial subsurface porosity and permeability. Where somesubsurface data are available, compaction models(Pittman and Larese, 1991; Gluyas and Cade, this vol-ume), the relationship of porosity vs. vitrinite reflectance(Schmoker and Gauthier, 1988; Schmoker and Hester,1990), or the predictive model of Scherer (1987) can beutilized for sandstones. If the prospective reservoir isexpected to be quartz rich (quartz arenite, subarkose,sublitharenite) the “Exemplar” model (Lander et al.,1995) can be an effective tool for predrill porosity evalua-tion (Lander and Walderhaug, 1997). “Exemplar” isbased on empirically calculated precipitation rates ofquartz cement in quartz-rich sandstones (Walderhaug,1994) ranging in age from Ordovician to Plio–Pleistocene(Lander et al., 1995). Significant progress in predictingquartz cementation rates with a minimum of basin-specific information has been recently reported byBjørkum et al. (in press). Each of these approaches has itslimitations and strengths and cannot be used indiscrimi-nantly. The applicability of some of these models toreservoir-quality assessment in frontier basins was dis-cussed by Bloch and Helmold (1995).

In mature basins, where calibration data sets areoften available, cement presence in the calibrationsamples is the determining factor in choosing the pre-dictive approach (Bloch and Helmold, 1995; Primmeret al., this volume). Weakly cemented sandstones dis-play “global” trends in reservoir quality, as first pub-lished by Scherer (1987). If cement in all or most of thesamples does not exceed 5–10%, multiple regressionanalysis can an effective predictive tool (Scherer, 1987;Bloch, 1991; Byrnes and Wilson, 1991). In uncementedor weakly cemented quartz-rich sandstones, the rela-tionship between porosity and effective stress derivedby Gluyas and Cade (this volume) can be very useful.Significant progress in prediction of reservoir qualityof quartz-poor sandstones was made by Wilson andByrnes (1988). Wilson and Byrnes generated a series of

proprietary linear regression functions for the predic-tion of porosity, permeability, and irreducible watersaturation in lithic sandstones. The functions werebased on a petrophysical and petrographic study of>500 samples representing a diverse suite of ductile-and volcanic-rich sandstones from various U.S. basins.Samples ranged in depth from 550 to 6460 m (1800 to21,200 ft) and in age from Early Cretaceous throughMiocene. The porosity function was able to predictporosity within a standard deviation of 1.9–2.2%.

Sandstones containing significant amounts ofcements appear to have predictable diagenetic styles(Primmer et al., this volume). In such sandstones, sev-eral scenarios exist for porosity prediction. In manyquartzose sandstones, quartz cementation is related todepth or burial history [e.g., Middle Cambrian sand-stones of the peri-Baltic area (Brangulis, 1985); Missis-sippian Kekiktuk sandstone of the North Slope ofAlaska (Bloch et al., 1990); Middle Jurassic sandstonesof the North Sea and Haltenbanken area offshore Nor-way (Bjørlykke et al., 1986, 1992; Bloch et al., 1986;Ehrenberg, 1990; Giles et al., 1992; Ramm, 1992; Wil-son, 1994]. Although many pay- and basin-specificpredictive relationships have been developed forquartzose sandstones, at this time only Exemplarappears to provide a more general predictive tool(Lander and Walderhaug, 1997).

Where cementation is not directly related to burialhistory, a satisfactory predictive model for sampleswith a wide range of cement content can be obtainedby grouping the data into two or more subsets anddeveloping a predictive model for each subset (Blochand Helmold, 1995). If controls on the distribution ofcement cannot be quantified, a qualitative (high-low)assessment is usually possible. Even in rocks with acomplex diagenetic history, reservoir quality is fre-quently related to simple parameters, such as grainsize (for a given provenance and burial history). Forexample, in the Norphlet Formation, stylolitization(not just intergranular pressure dissolution) andquartz cementation have been shown to be affected bygrain size (Thomas et al., 1993). As noted by Taylorand Soule (1993, p. 1554) for the North Bellridge field(California), “despite the important effects of diagene-sis, reservoir quality is still a function of the change ingrain size associated with depositional processes.”Usually the relationship of grain size and permeabilityis not expressed as a simple correlation. Rather, inmany reservoirs, sandstones coarser than a certaingrain size are characterized by permeabilities exceed-ing a cutoff value (Bloch and McGowen, 1994). Thisrelationship allows assessment of reservoir qualitybased on a facies model, assuming a depositionalfacies control of sand texture.

Future Trends

Although significant progress in reservoir qualityprediction has been made in the last decade, there isclearly a need for methodologies that are both moregeneral (“global”) and more accurate. The emphasis ofeffect-oriented/empirical modeling will be on expertsystems, hybrid process-effect approaches, nonlinear

Page 9: Reservoir Quality Prediction in Sand and Carbonates

multivariate regression analyses, possibility analysis,and neural networks (Wood and Byrnes, 1994).

Future activities in process-oriented/geochemicalmodeling will be focused on: (1) code development(recoding that makes programs “more user-friendly,more transportable between various operating sys-tems, and better suited to a modern coding environ-ment”), (2) improvement of mass transfer algorithms,and (3) development of a universal and robust, easilyupdatable database for minerals and aqueous species(Wood and Byrnes, 1994, p. 395). Most importantly,the quantitative effects of subsurface rock-fluid inter-action on porosity/permeability and the significanceof local vs. allochthonous cement sources need to bebetter understood.

Carbonates

In spite of the complexities of carbonate systems,advances in our ability to predict reservoir quality in advance of drilling have been made. Current suc-cesses, because of the complexities discussed above,have been with empirical approaches and three-dimensional reservoir models.

Process-Oriented Models

Process-oriented studies and models in carbonatesare very useful in our understanding of the mecha-nisms and complexities of aragonite, calcite, anddolomite precipitation and dissolution, and their inter-action with various diagenetic fluids. Back and Han-shaw (1971), Kharaka and Barnes (1973), Berner (1975),Parkhurst et al. (1980), Matthews and Froelich (1987),Banner and Hanson (1990), Dewers and Ortoleva(1990, 1994), Dreybrodt (1990), Quinn and Matthews(1990), and Kaufman (1994) have studied variousaspects of process-oriented modeling of carbonatesand diagenetic fluids. Most of the models calculategeochemical parameters of the water and rock duringreactions, without directly addressing changes inporosity and its distribution. Although these modelsprovide vast amounts of information and have fur-thered our understanding of carbonate diagenesis,because of the complexity of the chemical systems andbecause diagenetic environments change during pro-gressive burial of carbonates, none of these models caneffectively simulate reservoir quality evolution of shelflimestones or dolomites.

Empirical Models

Empirical techniques have been shown to be a pow-erful tool for the prediction of reservoir quality in car-bonates. Different approaches must be used dependingon the amount of subsurface data and whether outcropanalogs are present. In frontier areas, where analogousoutcrops are present, the methods of Tobin (this vol-ume), as discussed in the sandstone section, offer aviable technique to predict reservoir quality. Tobin usesexamples from both sandstone and carbonate outcrops.

In mature areas with extensive data sets, even giventhe potential for variability in preburial porosity,

empirical studies clearly document the decrease inporosity of carbonates with burial depth (Scholle,1977, 1978, 1981; Schmoker and Halley, 1982; Halleyand Schmoker, 1983; Schmoker and Hester, 1983;Schmoker, 1984; Schmoker et al., 1985; Amthor et al.,1994; Brown, this volume). These empirical studies canbe subdivided into two main groups: those of pelagiclimestones composed of low-Mg calcite; and lime-stones and dolomites interpreted to have beendeposited in shallow marine depositional environ-ments. The subdivision, as acknowledged byresearchers (e.g., Scholle, 1981), is mainly for reasonsof depositional complexity and diagenetic potential.Data from the low-Mg calcite pelagic limestones(Scholle, 1977, 1978, 1981) have simpler diagenetic his-tories and, as a result, have significantly less scatter inthe data than in shallow marine counterparts. Pelagiccarbonates are relatively stable, with no significantpreburial porosity modification, and more predictablefacies trends. The result is that changes in porosity inpelagic carbonates are most affected by mechanicaland chemical compaction during burial (Scholle, 1977,1978, 1981). Prediction of porosity requires the under-standing of the maximum burial depth and the pore-water chemistry (Scholle, 1977).

Scatter in the data from shallow marine carbonatesis interpreted as being due to early diagenetic varia-tions in preburial porosity (Halley and Schmoker, 1983;Schmoker, 1984; Schmoker et al., 1985), which suggeststhat specific predictions of reservoir porosity may notbe possible. These studies show that porosity is relatedto burial pressure, temperature and time, and lithol-ogy (limestone, dolomite, and shale content). Deposi-tional fabrics (e.g., mudstone, wackestone, packstone,grainstone) do not display significant differences inaverage porosity, even though they do differ in therange in porosity values (Brown, this volume).

A different approach is presented by Love et al. (thisvolume), using statistical methods in data-intensiveareas to allow the predrill prediction of reservoir quality.The authors analyze detailed geological data with aneural network predictive technique.

Additional examples of empirical predictions ofcarbonate reservoir quality are provided by integratedstudies using a combination of stratigraphy, structuralgeology, petrophysics, seismic reflection data, produc-tion data, and numerical methods. The predictionswere verified as successes or nonsuccesses by subse-quent drilling (Maureau and van Wijhe, 1979; Serna,1984; Beliveau and Payne, 1991). The strength of thesestudies is in the analysis of successes and failures.

Studies integrating geological and petrophysicaldata have proven very useful for reservoir characteri-zation and detailed infill drilling. By integratingdetailed analyses of depositional facies, facies tracts,sequence stratigraphy (especially at the parasequencelevel), diagenesis, pore types, porosity, permeability,capillary pressure, and saturation data, workers havebeen able to predict reservoir quality, reservoir perfor-mance, and bypassed pay. Studies include those byAufricht and Koepf, (1957), Keith and Pittman (1983),

xv

Page 10: Reservoir Quality Prediction in Sand and Carbonates

xvi

Bebout et al. (1987), Lucia and Conti (1987), Alger et al.(1989), Lucia et al. (1992a, b), Lucia (1993, 1995), Ker-ans et al. (1994), Martin et al. (1997), and Major andHoltz (this volume), among others. Incorporation ofdata into three-dimensional visualization modelsallows for reservoir quality prediction based on empir-ical correlations. Excellent examples of this methodol-ogy are presented Eisenberg et al. (1994), Kerans et al.(1994), Lucia et al. (1995), Tinker and Mruk (1995),Weber et al. (1995), and Tinker (1996).

Future Trends

Because of the complexity of carbonates (theirextensive postdepositional modification, pore types,and reservoir-quality distribution), empirical predic-tions appear to be the only feasible way to realisticallypredict predrill reservoir quality. Future studies ofpredrill reservoir-quality prediction in carbonates areexpected to continue to focus on the integration ofdetailed studies of subsurface cores and/or outcropanalog facies, detailed analysis of diagenesis, petro-physical analyses (particularly pore and pore-throat-type distribution, saturation, and capillary pressuredata), production data, fluid-flow modeling, andreservoir simulation. By using three-dimensionalmodeling, all detailed variables can be mapped priorto drilling. As mentioned above, examples of thismethodology are presented by Eisenberg et al. (1994),Kerans et al. (1994), Lucia et al. (1995), Tinker andMruk (1995), Weber et al. (1995), and Tinker (1996).However, future studies must also include substantia-tion by subsequent drilling, and discussions of suc-cesses and failures of reservoir quality prediction.

OVERVIEW OF MEMOIR

The Memoir consists of 17 chapters emphasizingeither reservoir-quality prediction techniques orexploration and exploitation case studies. Because ofthe diversity of papers, Table 1 is provided to help thereader gain an overview of the individual papers,including information on location, reservoir age, reser-voir mineralogy, stage in the exploration cycle, toolsused, and techniques used.

We have subdivided the chapters into two groups,those that address approaches to reservoir quality pre-diction and those that represent specific case studies.As a result, the chapters are not strictly subdivided by“sandstone” and “carbonate” examples. We hope thatthis approach serves to “cross-pollinate” ideas amongworkers in the field.

Approaches to Reservoir Quality Prediction

Tobin

Tobin shows how data obtained from sandstoneand carbonate outcrop exposures can be used to eval-uate subsurface porosity and permeability in poten-tial reservoirs. His approach, based on a systematicdecision-tree analysis, can be very useful in explorationrisk assessment, particularly in frontier basins with

limited or no subsurface information. Case studiesfrom China, Myanmar, and Turkey illustrate the pro-posed procedure.

Gluyas and Cade

Gluyas and Cade present a new equation for com-pactional porosity reduction as a function of depth foruncemented, clean, ductile-grain-poor sandstonesunder hydrostatic pressure. The equation is based onfield and experimental data. A modification of theequation relates porosity to effective stress, rather thanto depth, and thus can be used to predict porosity inoverpressured sands in which overpressure is rela-tively “early.” This technique provides a convenientway to predict porosity in uncemented sands or toprovide an upper limit on porosity in sandstonesexpected to contain authigenic cements. This tech-nique, tested against a global data set, has an accuracyof +2.5 porosity units at 95% confidence limits.

Brown

Brown addresses the influence of carbonate miner-alogy, fabric, and shale content on the rate of porosityloss with burial. Because of the availability of modernwell log suites, the Mississippian of the U.S. WillistonBasin is used as a study area. Porosity data obtained atconsistent intervals [10 ft (3 m)] help eliminate sam-pling bias, thus allowing an understanding of basin-scale porosity-loss mechanisms. Brown concludes thatporosity is selectively preserved in dolomites (vs.limestones) at similar burial conditions, and thatporosity decreases with increasing temperature.Cementation is a more important factor in loss of car-bonate porosity than is mechanical compaction.

Love, Strohmenger, Woronow, and Rockenbauch

Love et al. present a statistical approach to thepredrill prediction of reservoir quality. The authorsstress that this methodology can be applied to bothcarbonate and siliciclastic reservoirs, and illustratetheir techniques with a study of the Permian Zechsteincarbonates of the Southern Zechstein Basin of northernGermany. A three-dimensional distribution of reser-voir attributes is obtained by integrating geologicaldata (facies, mineralogy, porosity, permeability, welllogs, geochemistry) for 287 wells and applying a statis-tical analysis of these data. Because of the complexityof the spatial distribution of porosity and permeabil-ity, a neural network predictive technique is proven tobe more effective than linear regression.

Primmer, Cade, Evans, Gluyas, Hopkins, Oxtoby,Smalley, Warren, and Warden

Based on an analysis of a “global” data set, Primmeret al. conclude that chemical diagenesis impacts sand-stones through five predictable diagenetic “styles”: (1)quartz, commonly with lesser amounts of diageneticclays, and late ferroan carbonate; (2) clay minerals(illite or kaolinite) with lesser amounts of quartz (orzeolite) and late carbonate; (3) early grain-coatingclays that may inhibit quartz cementation duringdeeper burial; (4) early evaporite or carbonate

Page 11: Reservoir Quality Prediction in Sand and Carbonates

cements, and (5) zeolites, often in association withchlorite and/or smectite and late nonferroan carbon-ates.

The chemical diagenetic styles are a function ofdetrital mineralogy, depositional environments, andburial histories. Once the chemical diagenetic style ispredicted, a “most likely” value of cement abundancecan be estimated. This value is then subtracted fromporosity values obtained from compaction curves orequations (e.g., Gluyas and Cade, see above).

Sombra and Chang

Sombra and Chang emphasize the correlationbetween a parameter they term “the time-depthindex” (TDI) and porosity. The TDI-porosity relation-ship for three lithological types of reservoirs wasestablished for Upper Jurassic to Tertiary sandstonesof the Brazilian continental margin. Their approachinvolves (1) integration of the area enclosed betweenthe time-depth axes and the burial history curve of asandstone body (TDI) and (2) correlation of the inte-grated “TDI” with the porosity of the correspondingsandstone. The porosity of a lithologically similarsandstone can then be predicted prior to drilling ifinformation on its burial history TDI is available. Thistechnique can be useful when vitrinite reflectance dataare not available to calibrate the vitrinite reflectance-porosity relationship in formations in which such rela-tionship exists.

Evans, Cade, and Bryant

Evans et al. discuss permeability prediction basedon a combination of empirical and modeling tech-niques. This approach can be used in both frontier anddata-rich areas. The main difficulty in applying it isposed by the limitations in predicting variations ingeologic factors that are used to predict permeability.Evans et al. demonstrate that, provided the input dataare accurate, the permeability modeling techniquecommonly is able to predict permeability to withinhalf an order of magnitude.

Ehrlich, Bowers, Riggert, and Prince

Ehrlich et al. apply petrographic image analysis todetailed porosity analysis to equate porosity ele-ments to variations in permeability. This approachcan be used to predict the highest permeability possi-ble in a reservoir as a function of depth or basin loca-tion for a particular fabric. The concept is applied toinvestigations of Miocene sandstones of the SatunField in the Pattani basin (Gulf of Thailand) andUpper Carboniferous sandstones from the Cherokeebasin (Oklahoma).

Cabrera-Garzón, Arestad, Dagdelen, and Davis

Seismic reflection data from the Devonian Niskudolomites of Joffrey Field, Western Canada Sedimen-tary Basin, were used by Cabrera-Garzón et al. forreservoir quality prediction. Geostatistical simulationof porosity distribution within the field was obtainedthrough the analysis of P- and S-wave travel timesfrom multicomponent (3D, 3C) seismic reflection

data, integrated with porosity, permeability, and pet-rographic information from cores. Correlation ofporosity and Vp/Vs allows prediction of the three-dimensional distribution of porosity.

Zempolich and Hardie

Using the Jurassic of the Venetian Alps of Italy astheir study area, Zempolich and Hardie utilize detailedfield relationships, supplemented with geochemistry,to better understand and predict the geometries, distri-bution, timing and mechanism of formation of poten-tial dolomite reservoirs. They further use petrographyto constrain the evolution of reservoir-qualitydolomites. The authors conclude that reservoir-gradeporosity is initiated by the replacement of limestoneby dolomite, but that reservoir-grade permeability iscreated later, through the progressive recrystallizationof the replacement dolomite.

Case Studies

Gluyas and Witton

The diagenetic sequence encountered in Miocenesandstones by a wildcat well in the southern Red Seawas nearly identical to that predicted prior to drilling.However, predrill assessment of the abundance ofauthigenic cements was too conservative. Early halite,although expected, formed a “killer” cement thatplugged the entire porosity in the target sandstone.This work shows that with minimal data, reasonablyaccurate diagenetic predictions can be made.

Ramm, Forsberg, and JahrenHigh porosity (>20%) in deeply buried (>4000 m)

Upper Jurassic sandstones of the Norwegian CentralGraben is interpreted to have been preserved bymicroquartz coats. These coats inhibit precipitation ofpore-filling syntaxial quartz overgrowths duringdeeper burial. Microquartz appears to occur withinisochronous layers and has most likely been sourcedby syndepositional volcanic glass or sponge spicules.

GluyasUnlike Ramm et al., Gluyas attributes differences in

porosity in Upper Jurassic sandstones of the Norwe-gian Central Graben to the competition of quartzcementation and oil emplacement (“race for space”).High porosity at deep burial depths is interpreted tobe the result of retardation of quartz cementation bypetroleum emplacement rather than by the presence ofmicroquartz coats. This philosophy was used to pre-dict the porosity of the reservoir in a prospect a fewkilometers from existing data. Three porosity modelswere constructed to represent cases of cementationbefore, during, and after oil emplacement. The mostlikely outcome was predicted to be synchronouscementation and oil emplacement; thus, the porositywas estimated accordingly. Once drilled, the prospectwas found not to contain oil but water; however, thecore porosity of the sand was identical to that for themodel in which cementation predated oil emplace-ment. Perhaps the oil will arrive shortly!

xvii

Page 12: Reservoir Quality Prediction in Sand and Carbonates

xviii

Cavallo and Smosna

Cavallo and Smosna present a case study of a reser-voir at the development stage, the MississippianGreenbrier Limestone of the U.S. Appalachian Basin,West Virginia. This study integrates FormationMicroscanner (FMS) logs into an analysis and drillingprogram of an ooid shoal complex. By calibratingfacies characteristics with the log response and inte-grating dip information from the logs, the authorsillustrate reservoir quality prediction at the develop-ment scale.

Major and Holtz

Reservoir quality prediction at the development andreservoir management stages is presented by Majorand Holtz. This study of the Permian San Andres For-mation, West Texas (U.S.A.) Permian Basin illustratesthe importance of reservoir quality prediction in amature basin. Major and Holtz determine that flowunits are controlled by a combination of depositionalfacies and subsequent diagenetic alteration, and areable to quantify the amount of bypassed oil in bothlow-permeability and heterogeneous flow units.

Mountjoy and Marquez

Detailed petrographic studies of the DevonianLeduc Formation of the Western Canada SedimentaryBasin are presented by Mountjoy and Marquez. Reser-voir character of the dolomites is complex and can beobserved at different scales. The distribution of poretypes is controlled by original depositional facies,whereas the distribution of permeability is more afunction of diagenetic processes, especially dolomiti-zation. Mountjoy and Marquez compare dolomitesand limestones at variable burial depths, and illustratethat dolomites have higher porosity and permeabilitythan limestones at similar depths, because thedolomites are more resistant to pressure solution.

Smosna and Bruner

The content of shale and phyllite rock fragments inthe Devonian Lock Haven Formation of theAppalachian Basin (U.S.A.) is controlled by deposi-tional environments. The best reservoir quality occursin depositional facies characterized by an intermediatelabile grain content (distributary mouth bar and shelf).In those sandstones, secondary (lithmoldic) porosityenhances primary porosity. By contrast, sandstoneswith a low content of lithic grains (barrier island) havelow lithmoldic and total porosity. Porosity in sand-stones with a high abundance of lithic rock fragments(fluvial) was lost early due to compaction, thus pre-venting subsequent generation of lithmoldic porosity.

ACKNOWLEDGMENTS

We would like to extend our sincere thanks to thefollowing individuals who dedicated their time andeffort, and shared their expertise, toward improvingthe quality of the manuscripts in this Memoir: JohnAggatt (Lincolnshire, England), John Bell (Bogota,

Colombia), Mike Bowman (London, England),Andrew Brayshaw (Anchorage, Alaska, U.S.A.), SeanBrennan (Lawrence, Kansas, U.S.A.), Alton Brown(Plano, Texas, U.S.A.), Steve Bryant (Milan, Italy),Charles Curtis (Manchester, England), Martin Emery(Dallas, Texas, U.S.A.), Paul Enos (Lawrence, Kansas,U.S.A.), Laura Foulk (Denver, Colorado, U.S.A.),Steven Franks (Plano, Texas, U.S.A.), Mitch Harris (LaHabra, California, U.S.A.), Richard Heaton (Edin-burgh, Scotland), Andrew Horbury (London, Eng-land), Neil Hurley (Denver, Colorado, U.S.A.), KerryInman (Houston, Texas, U.S.A.), Nev Jones (Caracas,Venezuela), Marek Kacewicz (Plano, Texas, U.S.A.),Rob Kendall (Houston, Texas, U.S.A.), Andy Leonard(Aberdeen, Scotland), Bob Loucks (Plano, Texas,U.S.A.), Jerry Lucia (Austin, Texas, U.S.A.), Rick Major(Austin, Texas, U.S.A.), Jim Markello (Dallas, Texas,U.S.A.), Pascual Marquez (Maturin, Venezuela), Mal-colm McClure (London, England), Mark Osborne(Durham, England), Jackie Platt (London, England),David Roberts (London, England), Jim Schmoker(Denver, Colorado, U.S.A.), Per Svela (Stavanger, Nor-way), Dick Swarbrick (Durham, England), Pete Turner(Birmingham, England), and Bill Zempolich (Dallas,Texas, U.S.A.). The photomicrographs on the dustcover were taken by Mark Hopkins (London, Eng-land). Comments by Alton Brown, Dick Larese, MikeWilson, and Neil Hurley improved the introduction tothe Memoir. We also acknowledge the diligent workof the AAPG editorial staff, including Kevin Biddle,Neil Hurley, Ken Wolgemuth, and Anne Thomas.

REFERENCESAlger, R.P., D.L. Luffel, and R.B. Truman, 1989, New

unified method of integrating core capillary pressuredata with well logs: Society of Petroleum FormationEvaluation, v. 4, p. 145–152.

Amthor, J.E., E.W. Mountjoy, and H.G. Machel, 1994,Regional-scale porosity and permeability variationsin Upper Devonian Leduc buildups: implicationsfor reservoir development and prediction in car-bonates: AAPG Bulletin, v. 78, p. 1541–1559.

Aufricht, W.R., and E.H. Koepf, 1957, The interpreta-tion of capillary pressure data from carbonate reser-voirs: Transactions of the American Institute ofMining, Metallurgical, and Petroleum Engineers, v. 210, p. 402–405.

Back, W., and B.B. Hanshaw, 1971, Rates of physicaland chemical processes in a carbonate aquifer:Advances in Chemistry, v. 106, p. 77–93.

Banner, J.L., and G.N. Hanson, 1990, Calculation ofsimultaneous isotopic and trace element variationsduring water-rock interaction with applications tocarbonate diagenesis: Geochimica et CosmochimicaActa, v. 54, p. 3123–3137.

Bathurst, R.G.C., 1975, Carbonate sediments and theirdiagenesis: Developments in Sedimentology 12:New York, Elsevier, 658 p.

Bebout, D.G., and R.G. Loucks, eds., 1977, Cretaceouscarbonates of Texas and Mexico, applications to sub-surface exploration: University of Texas Bureau ofEconomic Geology Report of Investigations 89, 332 p.

Page 13: Reservoir Quality Prediction in Sand and Carbonates

Bebout, D.G., F.J. Lucia, C.F. Hocott, G.E. Fogg, andG.W. Vander Stoep, 1987, Characterization of theGrayburg reservoir, University Lands Dune field,Crane County, Texas: University of Texas at AustinBureau of Economic Geology Report of Investiga-tions 168, 98 p.

Beliveau, D., and D.A. Payne, 1991, Analysis of water-flood response of a naturally fractured reservoir:Society of Petroleum Engineers 22946, p. 603–613.

Berner, R.A., 1975, Diagenetic models of dissolvedspecies in the interstitial waters of compacting sedi-ments: American Journal of Science, v. 275, p. 88–96.

Berner, R.A., 1980, Early diagenesis: a theoreticalapproach: Princeton, New Jersey, Princeton Univer-sity Press, 241 p.

Bjørkum, A.A., and O. Walderhaug, 1990, Geometricalarrangement of calcite cementation within shallowmarine sandstones: Earth Science Reviews, v. 29, p. 145–161.

Bjørkum, P.A., E.H. Oelkers, P.N. Nadeau, O. Walder-haug, and W.M. Murphy, in press, Porosity predic-tion in quartz-rich sandstones as a function of time,temperature, depth, stylolite frequency, and the pres-ence of hydrocarbons: AAPG Bulletin, May, 1988.

Bjørlykke, K., P. Aaagard, H. Dypvik, D.S. Hastings,and A.S. Harper, 1986, Diagenesis and reservoirproperties of Jurassic sandstones from the Hal-tenbanken area, offshore mid-Norway, in A.M.Spencer, ed., Habitat of hydrocarbons on the Nor-wegian continental shelf: Norwegian PetroleumSociety, p. 275–286.

Bjørlykke, K., T. Nedkvitne, M. Ramm, and G.C. Saigal,1992, Diagenetic processes in the Brent Group (Mid-dle Jurassic) reservoirs of the North Sea: an overview,in A.C. Morton, R.S. Haszeldine, M.R. Giles, and S.Brown, eds., Geology of the Brent Group: GeologicalSociety Special Publication 61, p. 263–287.

Bloch, S., 1991, Empirical prediction of porosity andpermeability in sandstones: AAPG Bulletin, v. 75, p. 1145–1160.

Bloch, S., 1994, Secondary porosity in sandstones: sig-nificance, origin, relationship to subaerial unconfor-mities, and effect on predrill reservoir qualityprediction, in M.D. Wilson, ed., Reservoir qualityassessment and prediction in clastic rocks: SEPMShort Course 30, p. 137–159.

Bloch, S., and S.G. Franks, 1993, Preservation of shal-low plagioclase dissolution porosity during burialand aluminum mass balance: AAPG Bulletin, v. 77,p. 1488–1501.

Bloch, S., and K.P. Helmold, 1995, Approaches topredicting reservoir quality in sandstones: AAPGBulletin, v. 79, p. 97–115.

Bloch, S., and J.H. McGowen, 1994, Influence of depo-sitional environment on reservoir quality predic-tion, in M.D. Wilson, ed., Reservoir qualityassessment and prediction in clastic rocks: SEPMShort Course 30, p. 41–57.

Bloch, S., J.H. McGowen, J.R. Duncan, and D.W. Briz-zolara, 1990, Porosity prediction, prior to drilling, insandstones of the Kekiktuk Formation (Mississip-pian), North Slope of Alaska: AAPG Bulletin, v. 74,p. 1371–1385.

Bloch, S., R.K. Suchecki, J.R. Duncan, and K. Bjørlykke,1986, Porosity prediction in quartz-rich sandstones:Middle Jurassic, Haltenbanken area, offshore cen-tral Norway (abs.): AAPG Bulletin, v. 70, p. 567.

Brangulis, A.P., 1985, Vend i kembriy Latvii: strati-grafiya, litologiya i kollektorskiye svoystva (TheVendian and Cambrian of Latvia: stratigraphy,lithology, and reservoir quality) (in Russian): Riga,Department of Natural Gas of the USSR, 134 p.

Brown, A., this volume, Porosity variation in carbon-ates as a function of depth: Mississippian MadisonGroup, Williston Basin, in J.A. Kupecz, J. Gluyas,and S. Bloch, eds., Reservoir quality prediction insandstones and carbonates: AAPG Memoir 69, p. 29–46.

Bruton, C.J., 1985, Predicting mineral dissolution andcementation during burial: synthetic diageneticsequences (abs.): SEPM Gulf Coast Section ProgramWith Abstracts, v. 6, p. 2–3.

Budd, D.A., U. Hammes, and H.L. Vacher, 1993, Cal-cite cementation in the upper Floridan aquifer: amodern example for confined-aquifer cementationmodels?: Geology, v. 21, p. 33–36.

Budd, D.A., A.H. Saller, and P.M. Harris, eds., 1995,Unconformities and porosity in carbonate strata:AAPG Memoir 63, 313 p.

Byrnes, A.P., 1994, Empirical methods of reservoir qual-ity prediction, in M.D. Wilson, ed., Reservoir qualityassessment and prediction in clastic rocks: SEPMShort Course 30, p. 9–21.

Byrnes, A.P., and M.D. Wilson, 1991, Aspects of poros-ity prediction using multivariate linear regression(abs.): AAPG Bulletin, v. 75, p. 548.

Choquette, P.W., and L.C. Pray, 1970, Geologic nomen-clature and classification of porosity in sedimentarycarbonates: AAPG Bulletin, v. 54, p. 207–250.

Dewers, T., and P. Ortoleva, 1990, Interaction of reac-tion, mass transport, and rock deformation duringdiagenesis: mathematical modeling of intergranularpressure solution, stylolites, and differential com-paction/cementation, in I.D. Meshri and P.J. Ortol-eva, eds., Prediction of reservoir quality throughchemical modeling: AAPG Memoir 49, p. 147–160.

Dewers, T., and P. Ortoleva, 1994, Formation of stylo-lites, marl/limestone alternations, and dissolution(clay) seams by unstable chemical compaction ofargillaceous carbonates, in K.H. Wolf and G.V.Chilingarian, eds., Diagenesis IV: Elsevier, NewYork, Developments in Sedimentology 51, 155–216.

Dorobek, S.L., 1987, Petrography, geochemistry, andorigin of burial diagenetic facies, Siluro–DevonianHelderberg Group (carbonate rocks), CentralAppalachians: AAPG Bulletin, v. 71, p. 492–514.

Dreybrodt, W., 1990, The role of dissolution kinetics inthe development of karst aquifers in limestone: amodel simulation of karst evolution: Journal ofGeology, v. 98, p. 639–655.

Ebanks, W.J., 1990, Geology of the San Andres reservoir,Mallet lease, Slaughter field, Hockley County, Texas:implications for reservoir engineering projects, in D.G. Bebout and P.M. Harris, eds., Geologic andengineering approaches in evaluation of San

xix

Page 14: Reservoir Quality Prediction in Sand and Carbonates

xx

Andres/Grayburg hydrocarbon reservoirs—PermianBasin: University of Texas Bureau of Economic Geol-ogy Publication, p. 75–85.

Ehrenberg, S.N., 1990, Relationship between diagenesisand reservoir quality in sandstones of the Garn For-mation, Haltenbanken, mid-Norwegian continentalshelf: AAPG Bulletin, v. 74, p. 1538–1558.

Ehrlich, R., et al., this volume, Detecting permeabilitygradients in sandstone complexes—quantifying theeffect of diagenesis on fabric, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir 69,p. 103–114.

Eisenberg, R.A., P.M. Harris, C.W. Grant, D.J. Goggin,and F.J. Conner, 1994, Modeling reservoir hetero-geneity within outer ramp carbonate facies using anoutcrop analog, San Andres Formation of the Per-mian Basin: AAPG Bulletin, v. 78, p. 1337–1359.

Emery, D., K.J. Myers, and R. Young, 1990, Ancientsubaerial exposure and freshwater leaching insandstones: Geology 18, p. 1178–1181

Emery, D., and A.G. Robinson, eds., 1993, Inorganicgeochemistry: applications to petroleum geology:London, Blackwell Scientific Publications, 254 p.

Enos, P., and L.H. Sawatsky, 1981, Pore networks inHolocene carbonate sediments: Journal of Sedimen-tary Petrology, v. 51, p. 961–985.

Evans, J., C. Cade, and S. Bryant, this volume, A geo-logical approach to permeability prediction in clas-tic reservoirs, in J.A. Kupecz, J. Gluyas, and S.Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69, p. 91–102.

Giles, M.R., and J.D. Marshall, 1986, Constraints on thedevelopment of secondary porosity in the sub-surface: re-evaluation of process: Marine andPetroleum Geology 7, p. 378–397.

Giles, M.R., S. Stevenson, S.V. Martin, S.J.C. Cannon,P.J. Hamilton, J.D. Marshall, and G.M. Samways,1992, The reservoir properties and diagenesis of theBrent Group: a regional perspective, in AC. Morton,R.S. Haszeldine, M.R. Giles, and S. Brown, eds.,Geology of the Brent Group: Geological SocietySpecial Publication 61, p. 289–327.

Glasmann, J.R., R.A. Clark, S. Larter, N.A. Briedis, andP.D. Lundegard, 1989, Diagenesis and hydrocarbonaccumulation, Brent Sandstone (Jurassic), Bergenarea, North Sea: AAPG Bulletin, v. 73, p. 1341–1360.

Gluyas, J.G., this volume, Poroperm prediction forreserves growth exploration: Ula Trend, NorwegianNorth Sea, in J.A. Kupecz, J. Gluyas, and S. Bloch,eds., Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 201–210.

Gluyas, J., and C.A. Cade, this volume, Prediction ofporosity in compacted sands, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir69, p. 19–28.

Gluyas, J.G., and T. Witton, this volume, Poroperm pre-diction for wildcat exploration prospects: MioceneEpoch, Southern Red Sea, in J.A. Kupecz, J. Gluyas,and S. Bloch, eds., Reservoir quality prediction insandstones and carbonates: AAPG Memoir 69, p. 163–176.

Gluyas, J.G., A.G. Robinson, and T.P. Primmer, 1997,Rotliegend sandstone diagenesis: a tale of two waters,in J. Hendry, P. Carey, J. Parnell, A. Ruffel, and R.Worden, eds., Geofluids II 1997: Belfast, The Queen'sUniversity of Belfast, p. 291–294.

Grover, G., Jr., and F.J. Read, 1983, Paleoaquifer anddeep burial related cements defined by regionalcathodoluminescent patterns, Middle Ordoviciancarbonates, Virginia: AAPG Bulletin, v. 67, p. 1275–1303.

Halley, R.B., and D.K. Beach, 1979, Porosity preserva-tion and early freshwater diagenesis of marine car-bonate sands (abs.): AAPG Bulletin, v. 63, p. 460.

Halley, R.B., and J.W. Schmoker, 1983, High-porosityCenozoic carbonate rocks of South Florida: progres-sive loss of porosity with depth: AAPG Bulletin, v. 67, p. 191–200.

Handford, C.R., and R.G. Loucks, 1993, Carbonatedepositional sequences and systems tracts—responses of carbonate platforms to relative sea levelchanges, in R.G. Loucks and J.F. Sarg, eds., Carbon-ate sequence stratigraphy: recent developments andapplications: AAPG Memoir 57, p. 1–41.

Harrison, W.J., 1989, Modeling fluid/rock interactionsin sedimentary basins, in T. A. Cross, ed., Quantita-tive dynamic stratigraphy: New York, PrenticeHall, p. 195–231.

Harrison, W.J., and R.N. Tempel, 1993, Diageneticpathways in sedimentary basins, in A.D. Horburyand A.G. Robinson, eds., Diagenesis and basin devel-opment: AAPG Studies in Geology 36, p. 69–86.

Heald, M.T., and R.E. Larese, 1973, The significance ofthe solution of feldspar in porosity development:Journal of Sedimentary Petrology, v. 43, p. 458–460.

Hogg, A.J.C., P.J. Hamilton, and R.M. Macintyre,1993, Mapping diagenetic fluid flow within areservoir: K-Ar dating in the Alwyn area (UKNorth Sea): Marine and Petroleum Geology 10, p. 279–294.

James, N.P., 1978, Facies models: reefs: GeoscienceCanada, v. 5, p. 16–26.

Kaufman, J., 1994, Numerical models of fluid flow in car-bonate platforms: implications for dolomitization:Journal of Sedimentary Research, v. A64, p. 128–139.

Keith, B.D., and E.D. Pittman, 1983, Bimodal porosityin oolitic reservoir—effect on productivity and logresponse, Rodessa limestone (Lower Cretaceous),East Texas Basin: AAPG Bulletin, v. 67, p. 1391–1399.

Kerans, C., N.F. Hurley, and P.E. Playford, 1986,Marine diagenesis in Devonian reef complexes ofthe Canning Basin, western Australia, in J.H.Schroeder and B.H. Purser, eds., Reef diagenesis:New York, Springer-Verlag, p. 357–380.

Kerans, C., F.J. Lucia, and R.K. Senger, 1994, Inte-grated characterization of carbonate ramp reser-voirs using Permian San Andres Formation outcropanalogs: AAPG Bulletin, v. 78, p. 181–216.

Kharaka, Y.K., and I. Barnes, 1973, SOLMINEQ: asolution-mineral equilibrium computation: Spring-field, Virginia, National Technical Information Ser-vice Report PB 214-897, 82 p.

Page 15: Reservoir Quality Prediction in Sand and Carbonates

Kupecz, J.A., and L.S. Land, 1991, Late-stage dolomiti-zation of the Lower Ordovician Ellenburger Group,west Texas: Journal of Sedimentary Petrology, v. 61,p. 551–574.

Kupecz, J.A., and L.S. Land, 1994, Progressive recrys-tallization and stabilization of early-stage dolomite:Lower Ordovician Ellenburger Group, West Texas,in B. Purser, M. Tucker, and D. Zenger, eds.,Dolomites, a volume in honour of Dolomieu: IASSpecial Publication 21, p. 255–279.

Kupecz, J.A., I .P. Montañez, and G. Gao, 1993,Recrystallization of dolomite with time, in R.Rezak and D.L. Lavoie, eds., Carbonate microfabrics,frontiers in sedimentology: New York, Springer-Verlag, p. 187–194.

Land, L.S., 1980, The isotopic and trace element geo-chemistry of dolomite: the state of the art, in D.H.Zenger, J.B. Dunham, and R.L. Ethington, eds.,Concepts and models of dolomitization: SEPM Spe-cial Publication 28, p. 87–110.

Land, L.S., 1982, Introduction to dolomites anddolomitization: dolomites and dolomitizationschool: AAPG Course Notes, 29 p.

Land, L.S., 1985, The origin of massive dolomite: Jour-nal of Geological Education, v. 33, p. 112–125.

Land, L.S., 1986, Environments of limestone anddolomite diagenesis; some geochemical considera-tions, in J. Warme and K. Shanley, eds., Carbonatedepositional environments, modern and ancient,Part 5: diagenesis I: Colorado School of Mines Quar-terly, v. 81, no. 4, p. 26–41.

Lander, R.H., and O. Walderhaug, 1997, An empiri-cally calibrated model for sandstone reservoir qual-ity prediction (abs.): Program of the 1997 AnnualConvention of the AAPG, Dallas.

Lander, R.H., O. Walderhaug, A. Lyon, and A. Ander-sen, 1995, Reservoir quality prediction through sim-ulation of compaction and quartz cementation(abs.): Program of the 1995 Annual Convention ofthe AAPG, Houston, p. 53A.

Love, K.M., C. Strohmenger, A. Woronow, and K.Rockenbauch, this volume, Predicting reservoirquality using linear regression models and neuralnetworks, in J.A. Kupecz, J. Gluyas, and S. Bloch,eds., Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 47–60.

Lucia, F.J., 1993, Carbonate reservoir models: facies,diagenesis, and flow characterization, in D. Morton-Thompson and A.M. Woods, eds., Developmentgeology reference manual: AAPG Methods inExploration 10, p. 269–274.

Lucia, F.J., 1995, Lower Paleozoic cavern development,collapse, and dolomitization, Franklin Mountains,El Paso, Texas, in D.A. Budd, A.H. Saller, and P.M.Harris, eds., Unconformities and porosity in car-bonate strata: AAPG Memoir 63, p. 279–300.

Lucia, F.J., and R.D. Conti, 1987, Rock fabric, perme-ability, and log relationships in an upward-shoaling,vuggy carbonate sequence: University of Texas atAustin Bureau of Economic Geology GeologicalCircular 87-5, 22 p.

Lucia, F.J., C. Kerans, and R.K. Senger, 1992a, Definingflow units in dolomitized carbonate-ramp reser-voirs: Society of Petroleum Engineers, APE 24702,p. 399–406.

Lucia, F.J., C. Kerans, and G.W. Vander Stoep, 1992b,Characterization of a karsted, high-energy, ramp-margin carbonate reservoir: Taylor-Link West SanAndres unit, Pecos County, Texas: University ofTexas at Austin Bureau of Economic GeologyReport of Investigations 208, 46 p.

Lucia, F.J., C. Kerans and F.P. Wang, 1995, Fluid-flowcharacterization of dolomitized carbonate rampreservoirs: San Andres Formation (Permian) ofSeminole field and Algerita escarpment, PermianBasin, Texas and New Mexico, in E.L. Stoudt andP.M. Harris, eds., Hydrocarbon reservoir character-ization: SEPM Short Course 34, p. 129–153.

Major, R.P., and M.H. Holtz, this volume, Predictingreservoir quality at the development scale: methodsfor quantifying remaining hydrocarbon resource indiagenetically complex carbonate reservoirs, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 231–248.

Martin, A.J., S.T. Solomon, and D.J. Hartmann, 1997,Characterization of petrophysical flow units in car-bonate reservoirs: AAPG Bulletin, v. 81, p. 734–759.

Matthews, R.K., and X. Froelich, 1987, Forward model-ing of bank-margin carbonate diagenesis: Geology,v. 15, p. 673–676.

Maureau, G.T.F.R., and D.H. van Wijhe, 1979, The pre-diction of porosity in the Permian (Zechstein 2) car-bonate of eastern Netherlands using seismic data:Geophysics, v. 44, p. 1502–1517.

Meshri, I.D., 1989, On prediction of reservoir qualitythrough chemical modeling (abs.): AAPG Bulletin,v. 73, p. 390–391.

Meyers, W.J., 1991, Calcite cement stratigraphy: anoverview, in C.E. Barker and O.C. Kopp, eds.,Luminescence microscopy and spectroscopy: quali-tative and quantitative applications: SEPM ShortCourse 25, p. 133–148.

Meyers, W.J., and K.C. Lohmann, 1985, Isotope geo-chemistry of regionally extensive calcite cementzones and marine components in Mississippianlimestones, New Mexico, in N. Schneidermann andP.M. Harris, eds., Carbonate cements: SEPM SpecialPublication 36, p. 223–239.

Milliken, K.L., 1992, Chemical behavior of detritalfeldspars in mudrocks versus sandstones, Frio For-mation (Oligocene), South Texas: Journal of Sedi-mentary Petrology, v. 62, p. 790–801.

Milliken, K.L., E.F. McBride, and L.S. Land, 1989,Numerical assessment of dissolution versusreplacement in the subsurface destruction ofdetrital feldspars, Oligocene Frio Formation,south Texas: Journal of Sedimentary Petrology, v. 59, p. 740–757.

Mitchum, R.M., 1977, Seismic stratigraphy and globalchanges of sea level, part I: glossary of terms usedin seismic stratigraphy, in C.W. Payton, ed., Seismicstratigraphy applications to hydrocarbon explo-ration: AAPG Memoir 26, p. 205–212.

xxi

Page 16: Reservoir Quality Prediction in Sand and Carbonates

xxii

Mitchum, R.M., P.R. Vail, and S. Thompson III, 1977,Seismic stratigraphy and global changes of sealevel, Part II: the depositional sequence as a basicunit for stratigraphic analysis, in C.W. Payton, ed.,Seismic stratigraphy applications to hydrocarbonexploration: AAPG Memoir 26, p. 53–62.

Monty, C.L.V., D.W.J. Bosence, P.H. Bridges, and B.R.Pratt, 1995, Carbonate mud-mounds, their originand evolution: IAS Special Publication 23, 537 p.

Moore, C.H., 1989, Carbonate diagenesis and porosity:Developments in Sedimentology 46: New York,Elsevier, 338 p.

Morrow, D.W., 1982, Diagenesis 1. Dolomite—Part 2.Dolomitization models and ancient dolostones:Geoscience Canada, v. 9, p. 95–110.

Murray, R.C., 1960, Origin of porosity in carbonaterocks: Journal of Sedimentary Petrology, v. 30, p. 59–84.

Parkhurst, D.L., D.C. Thorstenson, and N. Plummer,1980, PHREEQUE: a computer program for geo-chemical calculations: USGS Water ResourcesInvestigational Report 80-96, 210 p.

Pittman, E.D., and R.E. Larese, 1991, Compaction oflithic sands: experimental results and applications:AAPG Bulletin, v. 75, p. 1279–1299.

Playford, P.E., 1980, Devonian “Great Barrier Reef” ofCanning Basin, western Australia: AAPG Bulletin,v. 64, p. 814–840.

Posamentier, H.W., M.T. Jervey, and P.R. Vail, 1988,Eustatic controls on clastic deposition I—sequencesand systems tracts models, in C.K. Wilgus, B.S.Hastings, C.G.St.C. Kendall, H.W. Posamentier,C.A. Ross, and J.C. Van Wagoner, eds., Sea-levelchanges: An integrated approach: SEPM SpecialPublication 42, p. 125–154.

Prezbindowksi, D.R., 1985, Burial cementation—is itimportant? A case study, Stuart City reef trend,south central Texas, in N. Schneidermann and P.M.Harris, eds., Carbonate cements: SEPM Special Pub-lication 36, p. 241–264.

Primmer, T.J., C.A. Cade, J. Evans, J.G. Gluyas, M.S.Hopkins, N.H. Oxtoby, P.C. Smalley, E.A. Warren,and R.H. Worden, this volume, Global patterns insandstone diagenesis: their application to reservoirquality prediction for petroleum exploration, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 61–78.

Quinn, T.M., and R.K. Matthews, 1990, Post-Miocenediagenetic and eustatic history of Enewetak Atoll:model and data comparison: Geology, v. 18, p. 942–945.

Ramm, M., 1992, Porosity-depth trends in reservoirsandstones: theoretical models related to Jurassicsandstones, offshore Norway: Marine and PetroleumGeology, v. 9, p. 563–567.

Ramm, M., A.W. Forsberg, and J.S. Jahren, this volume,Porosity-depth trends in deeply buried Upper Juras-sic reservoirs in the Norwegian Central Graben: anexample of porosity preservation beneath the nor-mal economic basement by grain-coating micro-quartz, in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,

Reservoir quality prediction in sandstones and car-bonates: AAPG Memoir 69, p. 177–200.

Robinson, A.G., and J.G. Gluyas, 1992, The duration ofquartz cementation in sandstones, North Sea andHaltenbanken basins: Marine and Petroleum Geol-ogy 9, p. 324–327

Rose, P.R., 1987, Dealing with risk and uncertainty inexploration: how can we improve?: AAPG Bulletin,v. 71, p. 1–16.

Saller, A.H., D.A. Budd, and P.M. Harris, 1994, Uncon-formities and porosity development in carbonatestrata: ideas from a Hedberg conference: AAPGBulletin, v. 78, p. 857–872.

Sarg, J.F., 1988, Carbonate sequence stratigraphy, inC.K. Wilgus, B.S. Hastings, C.G.St.C. Kendall, H.W.Posamentier, C.A. Ross, and J.C. Van Wagoner,eds., Sea-level changes: An integrated approach:SEPM Special Publication 42, p. 155–181.

Scherer, M., 1987, Parameters influencing porosity insandstones: a model for sandstone porosity predic-tion: AAPG Bulletin, v. 71, p. 485–491.

Schlager, W., 1992, Sedimentology and sequencestratigraphy of reefs and carbonate platforms:AAPG Continuing Education Course Note Series 34,71 p.

Schmoker, J.W., 1984, Empirical relation between car-bonate porosity and thermal maturity: an approachto regional porosity prediction: AAPG Bulletin, v. 68, p. 1697–1703.

Schmoker, J.W., and D.L. Gautier, 1988, Sandstoneporosity as a function of thermal maturity: Geology,v. 16, p. 1007–1010.

Schmoker, J.W., and R.B. Halley, 1982, Carbonateporosity vs. depth: a predictable relation for SouthFlorida: AAPG Bulletin, v. 66, p. 2561–2570.

Schmoker, J.W., and T. Hester, 1983, Porosity and ther-mal maturity of limestone bodies in Jurassic SwiftFormation, Williston Basin, North Dakota: U.S.Geological Society Open-File Report 83-723, 7 p.

Schmoker, J.W., and T.C. Hester, 1990, Regional trendsof sandstone porosity vs. vitrinite reflectance—apreliminary framework, in V.F. Nuccio and C.E.Barker, eds., Applications of thermal maturity stud-ies to energy exploration: Rocky Mountain Sectionof SEPM, p. 53–60.

Schmoker, J.W., K.B. Krystinik, and R.B. Halley, 1985,Selected characteristics of limestone and dolomitereservoirs in the United States: AAPG Bulletin, v. 69, p. 733–741.

Schneidermann, N., and P.M. Harris, eds., 1985, Car-bonate cements: SEPM Special Publication 36, 379 p.

Scholle, P.A., 1977, Chalk diagenesis and its relation topetroleum exploration: oil from chalks, a modernmiracle?: AAPG Bulletin, v. 61, p. 982–1009.

Scholle, P.A., 1978, Porosity prediction in shallow ver-sus deep water limestones—primary porositypreservation under burial conditions: SPE 7554.

Scholle, P.A., 1981, Porosity prediction in shallow vs.deepwater limestones: Journal of Petroleum Tech-nology, p. 2236–2242.

Page 17: Reservoir Quality Prediction in Sand and Carbonates

Scholle, P.A., D.G. Bebout, and C.H. Moore, eds., 1983,Carbonate depositional environments: AAPGMemoir 33, 708 p.

Scholle, P.A., and R.B. Halley, 1985, Burial diagenesis:out of sight, out of mind!, in N. Schneidermann andP.M. Harris, eds., Carbonate cements: SEPM SpecialPublication 36, p. 309–334.

Schroeder, J.H., and B.H. Purser, eds., 1986, Reef dia-genesis: New York, Springer-Verlag, 455 p.

Serna, M.J., 1984, Porosity prediction using amplitudemapping: case study of the Cretaceous M-2 Lime-stone Member of the Napo Formation, Ecuador: 4thMeeting of Petroleum Exploration in the Sub-Andean Basins, Bolivariano Symposium, Bogota,Colombia, Memoir V2, no. 29, 9 p.

Sluijk, D., and J.R. Parker, 1984, Comparison ofpredrilling predictions with postdrilling outcomes,using Shell’s prospect appraisal system (abs.):AAPG Bulletin, v. 68, p. 528.

Sneider, R.M., 1990, Introduction: reservoir descrip-tion of sandstones, in J.H. Barwis, J.G. McPherson,and J.R.J. Studlick, eds., Sandstone petroleum reser-voirs: New York, Springer-Verlag, p. 1–3.

Sombra, C.L., and H.K. Chang, this volume, Burial his-tory and porosity evolution of Brazilian UpperJurassic to Tertiary sandstone reservoirs, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 79–90.

Stoudt, E.L., and P.M. Harris, 1995, Hydrocarbonreservoir characterization: geologic framework andflow unit modeling: SEPM Short Course 34, 357 p.

Surdam, R.C., and L.J. Crossey, 1987, Integrated diage-netic modeling: a process-oriented approach forclastic systems: Annual Review of Earth and Plane-tary Sciences, v. 15, p. 141-170.

Taylor, T.R., and C.H. Soule, 1993, Reservoir character-ization and diagenesis of the Oligocene 64-zonesandstone, North Belridge field, Kern County, Cali-fornia: AAPG Bulletin, v. 77, p. 1549–1566.

Thomas, A.R., W.M. Dahl, C.M. Hall, and D. York, 1993,40Ar/39Ar analyses of authigenic muscovite, timing ofstylolitization, and implications for pressure solutionmechanisms: Jurassic Norphlet Formation, offshoreAlabama: Clays and Clay Minerals, v. 41, p. 269–279.

Tinker, S.W., 1996, Building the 3-D jigsaw puzzle:applications of sequence stratigraphy to 3-D reser-voir characterization, Permian Basin: AAPG Bul-letin, v. 80, p. 460–485.

Tinker, S.W., and D.H. Mruk, 1995, Reservoir charac-terization of a Permian giant: Yates field, WestTexas, in E.L. Stoudt and P.M. Harris, eds., Hydro-carbon reservoir characterization: SEPM ShortCourse 34, p. 51–128.

Tobin, R.C., this volume, Porosity prediction in fron-tier basins: a systematic approach to estimating sub-surface reservoir quality from outcrop samples, inJ.A. Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 1–18.

Toomey, D.F., 1981, European fossil reef models:SEPM Special Publication 30, 546 p.

Tyler, N., W.E. Galloway, C.M. Garrett, and T.E. Ewing,1984, Oil accumulation, production characteristics,and targets for additional oil recovery in major oilreservoirs of Texas: University of Texas Bureau ofEconomic Geology Circular 84-2, 31 p.

Vail P.R., R.M. Mitchum, and S. Thompson III, 1977,Seismic stratigraphy and global changes in sealevel, part 3: relative changes of sea level fromcoastal onlap, in C.W. Payton, ed., Seismic stratigra-phy applications to hydrocarbon exploration:AAPG Memoir 26, p. 63–97.

Van Wagoner J.C., R.M. Mitchum, K.L. Campion, andV.D. Rahmanian, 1990, Siliciclastic sequence stratig-raphy in well logs, cores, and outcrops: AAPGMethods in Exploration Series 7, 55 p.

Van Wagoner, J.C. , H.W. Posamentier, R.M. Mitchum,Jr., P.R. Vail, J.F. Sarg, T.S. Loutit, and J. Hardenbol,1988, An overview of the fundamentals of sequencestratigraphy and key definitions, in C.K. Wilgus,B.S. Hastings, C.G.St.C. Kendall, H.W. Posamentier,C.A. Ross, and J.C. Van Wagoner, eds., Sea-levelchanges: An integrated approach: SEPM SpecialPublication 42, p. 39–45.

Walderhaug, O., 1994, Precipitation rates for quartzcement in sandstones determined by fluid inclusionmicrothermometry and temperature-history model-ing: Journal of Sedimentary Research, Section A, p. 324–333.

Walls, R.A., and G. Burrowes, 1985, The role of cemen-tation in the diagenetic history of Devonian reefs,western Canada, in N. Schneidermann and P.M.Harris, eds., Carbonate cements: SEPM Special Pub-lication 36, p. 185–220.

Waples, D.W., and H. Kamata, 1993, Modeling poros-ity reduction as a series of chemical and physicalprocesses: in A.G. Doré et al., eds., Basin Modeling:Advances and Applications: Amsterdam, Elsevier,Norwegian Petroleum Society Special Publication 3,p. 303–320.

Waples, D.W., M. Suizu, and H. Kamata, 1992, The artof maturity modeling, part 2: alternative models andsensitivity analysis: AAPG Bulletin, v. 76, p. 47–66.

Weber, L.J., F.M. Wright, J.F. Sarg, E. Shaw, L.P. Har-man, J.B. Vanderhill, and D.A. Best, 1995, Reservoirdelineation and performance: applications ofsequence stratigraphy and integration of petro-physics and engineering data, Aneth Field, south-east Utah, U.S.A., in E.L. Stoudt and P.M. Harris,eds., Hydrocarbon reservoir characterization:SEPM Short Course 34, p. 1–29.

Williams, L.B., R.L. Hervig, and K. Bjørlykke, 1997,New evidence for the origin of quartz cements inhydrocarbon reservoirs revealed by oxygen isotopemicroanalysis: Geochimica et Cosmochimica Acta,v. 61, p. 2529–2538.

Wilson, J.L., 1975, Carbonate facies in geologic history:New York, Springer-Verlag, 471 p.

Wilson, M.D., 1994, Case history — Jurassic sand-stones, Viking Graben, North Sea, in M.D. Wilson,

xxiii

Page 18: Reservoir Quality Prediction in Sand and Carbonates

xxiv

ed., Reservoir quality assessment and prediction inclastic rocks: SEPM Short Course 30, p. 367–384.

Wilson, M.D., and A.P. Byrnes, 1988, Porosity predic-tion in lithic sandstones (unpublished report), 234 p.

Wood, J.R., and A.P. Byrnes, 1994, Alternate andemerging methodologies in geochemical andempirical modeling, in M.D. Wilson, ed., Reservoirquality assessment and prediction in clastic rocks,SEPM Short Course 30, p. 395–399.

Zempolich, W.G., and L.A. Hardie, this volume,Geometry of dolomite bodies within deep-waterresedimented oolite of the Middle Jurassic VajontLimestone, Venetian Alps, Italy: analogs forhydrocarbon reservoirs created through burialdolomitization, in J.A. Kupecz, J. Gluyas, and S.Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69, p. 127–162.

Page 19: Reservoir Quality Prediction in Sand and Carbonates

1

Chapter 1

Porosity Prediction in Frontier Basins:A Systematic Approach to Estimating

Subsurface Reservoir Quality from OutcropSamples

R.C. TobinAmoco Corporation

Houston, Texas, U.S.A.

ABSTRACT

In frontier basins where subsurface data are limited, or absent altogether,the study of reservoir rocks exposed in surface outcrops may be the domi-nant (or only available) means of predicting subsurface reservoir quality.This chapter provides a systematic, decision-tree–based procedure for using exist-ing tools and techniques to evaluate potential subsurface reservoir quality whenonly surface outcrops are available. This approach is applicable to both car-bonate and terrigenous clastic reservoirs. With this system, outcrop sam-ples are classified into one of ten lithofacies types whose reservoirproperties are codependent on common diagenetic or burial processes. Theclassification subdivides outcrop samples into either “tight” or “porous”lithofacies, depending on the measured porosity relative to economic mini-mum. “Tight” rocks include six end-member lithofacies that were eithercemented or compacted during burial, or were originally tight at the timeof deposition. “Porous” rocks include four lithofacies types that are catego-rized by original depositional fabric and the degree of alteration by recentsurface weathering. Risk assessment for each of the ten lithofacies typesusing existing geological tools and techniques is discussed, along withguidelines for estimating potential subsurface porosity and permeability.Case histories that illustrate the recommended process for assessing riskare described from China, Myanmar (Burma), and Turkey.

Tobin, R.C., 1997, Porosity prediction in frontierbasins: a systematic approach to estimating sub-surface reservoir quality from outcrop samples, inJ.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 1–18.

Page 20: Reservoir Quality Prediction in Sand and Carbonates

2 Tobin

INTRODUCTION

Geological analyses of hydrocarbon systems oftenrequire surface outcrop studies, particularly in frontierbasins where subsurface information is sparse. Outcropexposures provide the explorationist with uniqueopportunities for observing surface structural features,lateral bedding and facies variations, and three-dimen-sional spatial configurations that are less directlyobserved in the subsurface. This information is espe-cially beneficial in stratigraphic, sedimentologic, andstructural modeling studies. The generally unlimitedavailability of rock material also favors studies in whichlarge bulk samples must be used (e.g., source rock geo-chemistry, paleontology, geophysical rock properties).

Reservoir studies are also enhanced by outcrop inves-tigations. Reservoir facies predictions benefit greatly fromthe three-dimensional characteristics observable on out-crop exposures. Reservoir quality predictions benefit fromthe unlimited sample size availability and from thepotential for documenting the lateral variability inpetrophysical rock properties within a given facies.

Despite these advantages, outcrop-based predic-tions of subsurface reservoir quality are less direct (andpotentially less accurate) than those based on subsur-face data alone because of the following limitations.

1. Reservoir facies exposed at the surface may haveundergone a vastly different tectonic and burialhistory than their subsurface counterparts.

2. Diagenetic history and pore system evolution maybe different than that of subsurface counterparts.

3. Recent outcrop diagenesis (leaching, cementa-tion, sediment infill, etc.) may alter the composi-tion and pore system characteristics of ancientreservoir facies.

4. Basin-margin reservoir facies exposed on out-crops are less likely to contain hydrocarbonshows than their subsurface counterparts, andany shows that are present may be weatheredaway or severely biodegraded.

5. Outcrop exposures may be dominated by basin-margin reservoir facies rather than basin-centerfacies.

6. Reservoir rock provenance may vary from thesubsurface.

Therefore, interpretations of reservoir quality fromoutcrop data present a technical challenge to theexplorationist.

SCOPE AND INTENT

Various approaches to reservoir quality predictionfrom subsurface data are commonly used in explorationand are widely published. These predictions may rangefrom simple comparative analogs, where subsurfacedata are sparse, to more complex quantitative assess-ments of porosity and permeability from empirical cali-brations (Bloch, 1991; Byrnes, 1994; Bloch and Helmold,1995) or process-oriented geochemical models (Surdamand Crossey, 1987; Meshri, 1989; Meshri and Ortoleva,

1990; Wood, 1994). In contrast, surface outcrop studiesof potential reservoir facies are less commonly used inexploration, and are not as frequently documented inthe literature. With few exceptions (Goldstein, 1988;Scholle et al., 1991), most of the published outcrop stud-ies to date are only marginally related to the predictionof subsurface reservoir quality. Additionally, there is nopublished account of a deliberate, systematic approachto predicting subsurface reservoir quality from outcropsamples that explorationists can use to guide their stud-ies. As a result, outcrop evaluations are often plaguedby inefficiency and incomplete technical investigation.For example, potentially porous reservoir facies aresometimes overlooked as viable exploration targetsbecause of poor reservoir quality preservation at thesurface, even though the same facies may be highlyporous and permeable in the subsurface. Similarly, anunrealistically low degree of risk may be assigned to aporous facies exposed on outcrop whose original poresystem has been greatly enhanced by recent weatheringprocesses. The result of these problems may be poorrisk assessment of reservoir quality in frontier areas thatlack confirming subsurface well data.

To help alleviate these problems, this chapter isintended to provide a systematic approach for estimatingthe degree of risk associated with subsurface reservoirquality when only outcrop samples of prospective reser-voir facies are available for study. This approach uses adecision-tree–based process flow diagram to evaluate theuncertainties associated with reservoir quality usingexisting well-tested exploration tools and technologies(Table 1). As a by-product of this evaluation process, sur-face rock samples are classified into one of ten logicalgroupings (Types 1–10) whose predictabilities are code-pendent on common burial or diagenetic phenomena(Figure 1). The systematic approach described in thischapter can be used as a guide by explorationists chargedwith making either qualitative, quantitative, or semi-quantitative predictions in both carbonate and terrige-nous clastic strata in a time-effective manner. It is onlyintended to be a process-oriented approach to riskappraisal and is not intended as an all-inclusive solutionto subsurface porosity prediction. However, the effectiveapplication of this approach requires that the explo-rationist consider the wide variety of porosity construc-tive and destructive processes that commonly affectsubsurface reservoirs (Table 1). It also requires that theexplorationist utilize information from a variety ofrelated technologies in an integrated manner (Table 2).

RECOMMENDED APPROACH

The first recommended step in subsurface porosityprediction from surface data is to classify outcrop rocksamples into one of ten logical categories whosepresent-day porosity and permeability values are the endproducts of common geologic and burial phenomena(Table 3). This classification is organized by adecision-tree flow diagram that leads the explorationistto discover the cause(s) for the present-day porosity andpermeability of the rocks being studied (Figure 1). Theclassification then poses additional questions that require

Page 21: Reservoir Quality Prediction in Sand and Carbonates

a more detailed technical evaluation designed to assistthe explorationist in predicting reservoir quality andassigning an appropriate degree of exploration risk. Thefollowing discussion will lead the reader through a num-ber of key decision points illustrated in the decision treeshown in Figure 1.

First Decision: Porous or Tight?

The first decision point considered is whether the rocksample in question is currently “porous” or “tight” (Fig-ure 1). This distinction does not rely on a universal poros-ity cutoff, but is dependent on the economic requirementsof each play, and should include both the amount andtype of porosity present. For example, a thin, deeplyburied reservoir sandstone in a frontier area having smallstructural traps and lacking needed economic infrastruc-ture (pipelines, transportation, etc.) might require 20–25%porosity and >500 md of permeability for the play to beeconomic (i.e., to ensure economically viable reserves andassociated flow rates). In contrast, a thick, shallow, frac-tured carbonate play in an area having larger structuraltraps adjacent to an accessible pipeline might require only5% porosity to be economic. In either case, if a large pro-portion of the pore system has ineffective microporosity,then the porosity cutoff used would have to be consider-ably higher (Figure 2). Therefore, when establishing localporosity cutoffs, it is important that the explorationist useonly effective porosities. It is also critical to recognize thattight lithofacies exposed at the surface may have porouscounterparts in the subsurface, and vice versa.

TIGHT ROCKSDecision: Why Is the Sample Tight?

If the outcrop sample in question is deemed to be tight(effective porosity and permeability are below economicrequirements), the next decision point to be considered is“Why is the sample tight?” Three possibilities exist. Eitherthe original depositional fabric of the sample was tight,the original fabric has been so obscured by diagenesis thatthe cause of low porosity is uncertain, or the original sed-imentary fabric was porous, but postdepositional diagen-esis has reduced porosity to unacceptably low amounts.Each of these possibilities is described below.

Tight Depositional Facies (Rock Type 1)If it is determined from petrographic observations

that the original fabric was tight (i.e., a depositionalfacies lacking high initial macroporosity and perme-ability), the sample is considered to be a Type 1 lithol-ogy (Figure 1). For carbonate rocks, common examplesof Type 1 lithofacies would be marl, lime mudstone orwackestone, sandy lime mudstone to micritic sand-stone, or fine, dense crystalline dolomite. For clasticrocks, Type 1 lithofacies are usually shales or argilla-ceous siltstones, sandstones, and conglomerates(wackestones). Such lithofacies are usually associatedwith turbid, low-energy environments of depositionthat are not conducive to the deposition of sedimentswith high initial macroporosity. Hence, the explorationrisks associated with Type 1 lithofacies are normallyconsidered to be quite high, unless paleogeographic

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 3

Table 1. Geologic Factors Affecting Subsurface Reservoir Quality.

Geologic Factor Effect on Porosity

Ancient destructive diagenesis Reduces porosity(sediment infill, cementation, recrystallization, mechanical and chemical compaction)

Ancient constructive diagenesis Enhances porosity(dissolution, fracturing)

Other ancient diagenesis May reduce or enhance porosity(mineralogic replacement, authigenic clay growth, brecciation, tectonic deformation)

Framework composition/provenance May control postdepositional diagenesisEnvironment of deposition Controls prediagenesis porosityPaleoclimate Affects EOD, weathering, karstificationDepth of burial Indirectly related to porosity lossPressuring and overpressuring Early overpressure may enhance porosity

Thermal maturity Indirectly related to porosity lossErosional events/unconformities May reduce or enhance porosityPore fluid migration (water) Enhances cementation/dissolutionPore fluid migration (oil) Inhibits cementation/dissolutionAssociated rock strata—seal Affects pore fluid entrapmentAssociated rock strata— source rock Controls type of migrating pore fluids

Page 22: Reservoir Quality Prediction in Sand and Carbonates

4 Tobin

reconstructions can be used to predict the occurrence ofmore porous depositional facies elsewhere.

Lower risks can also be predicted for Type 1 litho-facies under the following conditions: (1) Fracturing ofbrittle rock strata can create economic reservoirs out ofnonporous or low-porosity facies, although the domi-nant effect is to significantly increase permeability ratherthan porosity. The likelihood of fracture development isrelated to structural position and various rock properties.In general, closely spaced fracture networks are favoredby rocks with low matrix porosities that are fine grained,thinly bedded, and composed predominantly of brittle(nonductile) minerals such as quartz, feldspar, dolomite,and calcite (Nelson, 1985). (2) Karstification of tight car-bonate facies can create economic porosity, but the pre-diction of porosity in karsted facies is severely limited bythe lack of adequate diagenetic models (Saller et al.,1994). Total porosity in karstified reservoirs is generallylow (3–6%), with associated low permeability exceptwhere fractured and leached. Karstic reservoirs may alsobe highly compartmentalized (Tobin, 1985; Kerans,1988). (3) Dolomitization has been reported to creatematrix porosity in originally tight, micritic facies (Weyl,1960), although more recent evidence suggests thatdolomitization alone may be insufficient to create aviable pore system. Mineralogically selective, post-dolomite dissolution, for example, may be necessary(Ottmann et al., 1976; Lucia and Major, 1994; Purser et al.,1994). (4) Subsurface (burial) dissolution may also createeconomic porosity, but only if there are dissolvable con-stituents within the rock and a sufficient plumbing sys-tem existed for subsurface dissolution to have beeneffective. For each of these four mechanisms, burial his-tory reconstructions (time-temperature-depth profiles)may offer clues to past episodes in which porosity-constructive events could have taken place. Burial histories

and associated paragenetic sequences should always beincluded in risk assessment of Type 1 lithofacies beforedisregarding them as prospective reservoirs (Figure 3).

Uncertain Depositional Facies (Rock Type 2)

For some rock samples, the distinction between tightfacies and tight diagenesis is uncertain because the orig-inal depositional fabric has been obscured by diagenesis(ancient or recent). These are referred to as Type 2 rocks(Figure 1). Examples of Type 2 rocks are recrystallizedsparry limestones, some dolomites (particularly coarsecrystalline, nonplanar dolomites of burial origin), andsome quartzose sandstones whose original fabrics havebeen obscured by intense quartz cementation and asso-ciated pressure solution or incipient metamorphism.For these lithofacies, subsurface porosity prediction isuncertain, and exploration risk is considered to be quitehigh unless fracturing, karsting, dolomitization, or dis-solution can be predicted elsewhere. For the explo-rationist faced with this type of facies, the recognition oforiginal sedimentary fabric is of paramount impor-tance. Tools that may be used to help identify these fab-rics are “white-card” transmitted light microscopy(Folk, 1987), thin-section epifluorescence (Dravis andYurewicz, 1985), and cathodoluminescence (Figure 4).

Destructive Diagenesis (Rock Types 3–6)

Some potential reservoir rocks exposed at the sur-face originally had porous depositional fabrics, butwere subsequently altered by porosity-destructivediagenetic processes. Typical carbonate lithofacies inthis group include grainstones, packstones, some pla-nar dolomites, and some reefal facies. Typical clasticrepresentatives are clean, matrix-poor sandstones andconglomerates (arenites). Risk assessment for this

Why isit tight?

Tight F

acies

Destructive

Diagenesis

TYPE 1

When diddiagenesisdestroyporosity?

TYPE 2

TYPE 3LowRisk

Recent

Ancient

What wasthedominantmechanism?

Compac

tion TYPE 4

TYPE 5

TYPE 6

HighRisk

Mod-HighRisk

HighRisk

HighRisk

Uncertain

LateCements

Early

Cements

Mod-HighRisk

How muchsurfaceweathering?

Was theoriginalfabricporous ortight?

Minim

al

Uncertain

PorousDominant

TYPE 7

TYPE 10UncertainRisk

Tight

TYPE 8

TYPE 9HighRisk

LowRisk

Is thesampleporousor tight?

Porous

Tigh

t

Mod-HighRisk

Figure 1. Decision-tree flowdiagram used to evaluate thedegree of exploration riskfrom the study of outcropmaterials. Outcrop samplesare classified into ten rocktype categories based ontheir potential as subsurfacereservoir facies. Completedescriptions of each rocktype are given in the textand are summarized inTable 3.

Page 23: Reservoir Quality Prediction in Sand and Carbonates

group requires that the timing and type of destructivediagenetic processes be identified, leading to the nextdecision points in the decision tree shown in Figure 1.

Decision: When Did Diagenesis Destroy Porosity?

For facies having initially porous depositional fabrics,the next question to consider is “When did the porosity-destructive diagenesis occur?” Two basic possibilitiesexist. Either the porosity was destroyed during recentoutcrop exposure or it was destroyed during ancientdiagenetic event(s). Both possibilities are discussedbelow.

Recent Pore Destruction (Rock Type 3)

Some outcrop samples show clear evidence ofrecent outcrop-related pore destruction, and are con-sidered Type 3 lithofacies (Figure 5). Type 3 rocks orig-inally contained economic amounts of primary,secondary, or dissolution-enhanced primary porositythat survived burial diagenesis prior to recent outcropexposure. Upon exposure, the pore system of theserocks was subsequently destroyed by a variety of sur-face and near-surface diagenetic processes. Theseinclude recent sediment infill (e.g., terra rosa or othersoil-forming processes), infill by weathering by-prod-ucts such as iron oxide or clays, oil biodegradationresulting in pore-plugging bitumen, or surface to near-surface cementation.

Outcrop-related cements may be difficult to distin-guish from ancient cements, although some petro-graphic criteria exist for their recognition. These include

the presence of pendant or meniscus morphologies (par-ticularly if they follow obvious by-products of burialdiagenesis), isotopic and/or trace element compositionscharacteristic of meteoric origin and unrelated to priordiagenetic byproducts, and fluid inclusions suggestiveof recent exposure (e.g., air inclusions and/or domi-nantly monophase aqueous inclusions, particularly iftwo-phase inclusions are present in earlier cement gen-erations). The exploration risks associated with Type 3(recent pore destruction) reservoir facies are consideredto be relatively low because such facies clearly containedeconomic porosity prior to outcrop exposure. Therefore,porous counterparts probably exist somewhere in thesubsurface, although they may not be ubiquitous. Forthe explorationist who wants to quantify the risk associ-ated with Type 3 reservoirs, the following questionsshould be thoroughly investigated: (1) How muchporosity was present in the sample prior to outcropexposure? (2) Was the porosity of the sample well con-nected (permeable) prior to outcrop exposure? (3) Howdeep was this sample buried prior to outcrop exposure?(4) How much deeper could economic porosity survivein this sample beyond its estimated pre-outcrop depth ofburial? (5) Are compaction-inhibiting processes (e.g.,early overpressuring) or compaction-enhancingprocesses (e.g., pressure solution) likely to affect theporosity vs. depth estimates? (6) Is the pre-outcropporosity primary or secondary? If it is primary, is thereany potential for porosity enhancement by secondarydissolution elsewhere? These questions should lead theexplorationist to a reasonable estimate of the range ofporosity and permeability likely to be encountered atany given drilling depth. An example of this process isillustrated in Table 4.

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 5

Table 2. Related Technologies Used in Porosity Prediction.

Technology Why Is It Important?

Sedimentology Facies analysis, environment of depositionPetrography Microfacies, diagenesis, pore system descriptionFluorescence Depositional/diagenetic fabric recognition,

pore geometryLuminescence Depositional/diagenetic fabric recognitionGeophysics Facies analysis, unconformity recognitionPaleontology Stratigraphy, environment of deposition,

unconformity recognitionCore analysis Porosity, permeability, pore geometryInorganic geochemistry Diagenetic interpretations, unconformity

recognitionOrganic geochemistry Source rock quality, migration timingFluid inclusion thermometry Migration timing, diagenesis, thermal maturityThermal maturity analyses Indirectly related to porosity, hydrocarbon

phase preservedBasin modeling Timing of porosity creation/destruction events,

depth of burialCompaction simulation Prediction of past burial depths and depth to

porosity basementRock mechanics Probability of fracturing

Page 24: Reservoir Quality Prediction in Sand and Carbonates

6 Tobin

Tab

le 3

. Su

mm

ary

of O

utc

rop

Cat

egor

ies

and

Ass

ocia

ted

Res

ervo

ir Q

ual

ity

Ris

ks.

Cat

egor

yN

ame

Por

osit

yT

ypic

al L

ith

olog

ies

Ris

k A

sses

smen

t

Typ

e 1

Tig

ht d

epos

itio

nal

Tig

htM

icri

tic

limes

tone

, mar

l and

sha

le, s

and

yH

igh

risk

unl

ess

frac

turi

ng, d

olom

itiz

atio

n,

faci

eslim

esto

ne, m

icri

tic

dol

omit

e, a

rgill

aceo

us

or p

oros

ity

can

be p

red

icte

d

silt

ston

e, s

and

ston

e, o

r co

nglo

mer

ate

Typ

e 2

Unc

erta

in d

epos

itio

nal

Tig

htR

ecry

stal

lized

spa

rry

limes

tone

, som

e co

arse

,H

igh

risk

as

abov

e, u

nles

s or

igin

al fa

bric

can

be

faci

esno

npla

nar

dol

omit

e, s

ome

quar

tz-c

emen

ted

det

erm

ined

or

met

amor

phos

ed q

uart

z sa

ndst

ones

Typ

e 3

Rec

ent p

ore

des

truc

tion

Tig

htO

rigi

nally

por

ous

sand

ston

es a

nd c

ongl

omer

ates

, V

ery

low

ris

k fo

r pr

ospe

cts

shal

low

er th

an

lime

grai

nsto

nes

or p

acks

tone

s ti

ghtl

y ce

men

ted

pre-

outc

rop

buri

al d

epth

; var

iabl

e ri

sk fo

r by

rec

ent w

eath

erin

g by

-pro

duc

ts

dee

per

pros

pect

s

Typ

e 4

Dom

inan

tly

com

pact

edT

ight

Ori

gina

lly p

orou

s, b

ut n

ow ti

ghtl

y co

mpa

cted

Ver

y hi

gh r

isk

for

pros

pect

s th

at a

re a

s d

eep

sand

ston

es, c

ongl

omer

ates

, or

nonm

icri

tic

as p

re-o

utcr

op b

uria

l dep

th u

nles

s ea

rly

carb

onat

esov

erpr

essu

ring

, rim

cem

enta

tion

, or

dis

solu

tion

can

be

pred

icte

d

Typ

e 5

Ear

ly n

ear-

surf

ace

Tig

ht

Ori

gina

lly p

orou

s sa

ndst

ones

and

con

glom

erat

es,

Hig

h ri

sk u

nles

s la

tera

l cem

ent p

inch

out o

rce

men

ted

lime

grai

nsto

nes,

or

pack

ston

es ti

ghtl

y ce

men

ted

cem

ent d

isso

luti

on c

an b

e pr

edic

ted

by a

ncie

nt n

ear-

surf

ace

cem

ents

Typ

e 6

Lat

e bu

rial

cem

ente

dT

ight

Ori

gina

lly p

orou

s sa

ndst

ones

and

con

glom

erat

es,

Mod

erat

e to

hig

h ri

sk u

nles

s la

tera

l pin

chou

t, lim

e gr

ains

tone

s, o

r pa

ckst

ones

tigh

tly

cem

ente

ddi

ssol

utio

n, o

r dia

gene

tic tr

aps

can

be p

redi

cted

by a

ncie

nt b

uria

l cem

ents

Typ

e 7

Rec

ent w

eath

erin

gPo

rous

Any

por

ous

litho

logy

who

se p

ore

syst

em is

V

ery

low

ris

k fo

r pr

ospe

cts

shal

low

er th

an

min

imal

inhe

rite

d fr

om th

e su

bsur

face

(min

imal

rec

ent

pre-

outc

rop

buri

al d

epth

; var

iabl

e w

eath

erin

g)ri

sk fo

r d

eepe

r pr

ospe

cts

Typ

e 8

Wea

ther

ed;

Poro

usO

rigi

nally

por

ous

dep

osit

iona

l fab

rics

ren

der

ed

Mod

erat

e to

hig

h ri

sk fa

cies

; ris

k as

sess

men

t d

epos

itio

nal f

abri

cti

ght b

y co

mpa

ctio

n or

cem

enta

tion

, but

eq

uiva

lent

to T

ype

4, 5

, or

6 as

app

ropr

iate

po

rous

leac

hed

by

rece

nt w

eath

erin

g

Typ

e 9

Wea

ther

ed;

Poro

usO

rigi

nally

tigh

t dep

osit

iona

l fab

rics

that

hav

e be

enH

igh

risk

faci

es; r

isk

asse

ssm

ent e

quiv

alen

t d

epos

itio

nal f

abri

cle

ache

d b

y re

cent

wea

ther

ing

proc

esse

sto

Typ

e 1

litho

faci

es

tigh

t

Typ

e 10

Rec

ent w

eath

erin

gPo

rous

Any

res

ervo

ir li

thol

ogy

who

se p

ore

syst

em

Unc

erta

in r

isk,

but

gen

eral

ly h

ighe

r fo

r un

cert

ain

cont

ains

app

reci

able

am

ount

s of

sec

ond

ary

in

crea

sing

sec

ond

ary

poro

sity

com

pone

ntpo

rosi

ty o

f unc

erta

in o

rigi

n

Page 25: Reservoir Quality Prediction in Sand and Carbonates

Ancient Pore Destruction (Rock Types 4–6)

Most diagenetically altered outcrop samples showunmistakable evidence of ancient pore destructionresulting from some combination of compaction andcementation. Risk assessment for these rocks requires

that the dominant diagenetic mechanism for poredestruction be identified, leading to the next decisionpoint on the decision tree shown in Figure 1.

Decision: Ancient Pore Destruction—What Was the Dominant Mechanism?

Outcrop samples whose pore system was destroyedprimarily by burial compaction are referred to as Type 4(compacted) rocks. Cement-dominated samples arereferred to as either Type 5 (early near-surface cemented)or Type 6 (late burial cemented) rocks depending on thetiming of cement emplacement (Figure 1). In general, allthree rock types are high-risk lithologies, although thespecific degree of risk can be highly variable, dependingon cement type, volumetric amount, timing, temperature,and presence of hydrocarbons. For this category, theexplorationist must first determine the dominant mecha-nism of porosity loss before assessing exploration risk.

Compaction (Rock Type 4)

Type 4 lithofacies include nonargillaceous sandstonesand conglomerates (arenites) and nonmicritic, grain-supported carbonates (lime grainstones, some pack-stones, some reef rocks, and some dolomites) whosepore systems have been destroyed by either mechanicalcompaction (grain rotation, slippage, rearrangement,repacking, plastic deformation, or grain breakage), pres-sure solution (intergranular or whole rock), or both (Fig-ure 6). Type 4 rocks can contain minor amounts ofcement, but the dominant mechanism for porosity loss isfrom compaction. Associated intergranular volumes arecharacteristically low. Although mechanical compactioncan severely reduce porosity for sandstones of any com-position, it is most effective in those containing an abun-dance of ductile lithic grains (Pittman and Larese, 1991).Similarly, mechanical compaction is most effective ingrain-supported carbonate rocks containing ductilemicritic grains (peloids, onkoids, some intraclasts) ratherthan hard, brittle grains like ooids or bioclasts (Moore,1989, his figure 9.5). Compaction from pressure solutionis most effective in sandstones containing an abundanceof quartz and feldspar with minimal lithics (Pittman andLarese, 1991). Pressure solution is most likely to occur ingrain-supported carbonate rocks that contain metastable(aragonitic) grain types (Wagner and Matthews, 1982) orinsoluble components such as clays, quartz, and organ-ics (Weyl, 1959). The presence of oil within poresappears to retard the effects of pressure solution (Dun-nington, 1967; Feazel and Schatzinger, 1985).

Exploration risk for Type 4 reservoir facies is veryhigh for prospects that are as deep or deeper than pre-outcrop burial depth. Risk can be considerably lower,however, under any of the following conditions: (1)early, shallow overpressuring can retard the rate ofporosity loss from compaction (Scherer, 1987; Pittmanand Larese, 1991). (2) Risk can also be lower if it can bedemonstrated that early compaction-retarding rimcements are likely elsewhere. For example, early incipi-ent quartz overgrowth cement in sandstones (Pittmanand Larese, 1991) or early calcite rim cements or replace-ment dolomite in carbonate grainstones (Purser, 1978;

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 7

(A)

(B)

Figure 2. Porous reservoir facies containing apprecia-ble amounts of ineffective microporosity. (A) Ooid-skeletal lime grainstone from the oil-producing ArabZones in Dukhan field, Qatar, containing 20% totalporosity, 40% of which is ineffective microporosity(purple color, note arrow) found within micriticgrains. Total effective porosity for this sample isonly 12%, and includes both primary (P) and second-ary grain-moldic (M) pores (thin-section photomicro-graph using plane-transmitted and ultravioletfluorescent light; 80×). (B) Litharenite sandstonefrom western Siberia containing 17% total porosity,35% of which is ineffective microporosity associatedwith authigenic clays, mostly pore-filling kaolinite(K). Total effective porosity for this sample is only11% (SEM photomicrograph taken at 1400×).

Page 26: Reservoir Quality Prediction in Sand and Carbonates

8 Tobin

Moore and Druckman, 1981) can produce a rigid frame-work that is resistant to further compaction. Althoughsmall in volume, such cements take on the bulk of theoverburden pressure, thereby inhibiting grain slip-page/rotation, ductile grain deformation, and pressuresolution. (3) Dissolution of various unstable rock compo-nents (cements and grains) can create secondary porositythat could yield an economic pore system, although theprobability of significant porosity increase from the dis-solution of a tightly compacted rock is fairly low becauserocks generally have a very poor plumbing system forcirculating undersaturated fluids.

Early Near-Surface Cementation (Rock Type 5)Type 5 lithofacies are originally porous rocks that

have been tightly cemented during early diagenesis bya variety of surface, near-surface, and shallow burialcements (Figure 7). Intergranular volumes found inType 5 lithofacies are generally high because of thelimited amount of compaction associated with shallowburial. For carbonate rocks, typical cements includecalcite or aragonite of vadose, meteoric phreatic,marine or shallow burial origin, early dolomite, orevaporitic cements like anhydrite, gypsum, or halite.For sandstones, the most common early cements are

050100150200250300350400140

130

120

110

100

90

80

70

60

50

40

30

Tem

per

atu

re (

°C)

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Dep

th (

km)

unconformity (potential karst Ø)

thrusting & unroofing(potential fracture Ø)

oil generation / migration(potential dissolution Ø)

A

050100150200250300350400140

130

120

110

100

90

80

70

60

50

40

30

Time (Ma)

Tem

per

atu

re (

°C)

3.0

2.5

2.0

1.5

1.0

0.5

0.0

Dep

th (

km)late erosional unroofing

(minor potential forfracture Ø)

BTime (Ma)

Figure 3. Two hypothetical burialhistory plots of a high risk, Type 1reservoir facies (wrong deposition-al facies). In (A) two opportunitiesexisted for secondary porositydevelopment prior to late thrustingand possible fracture porosityenhancement. In (B) little or noopportunity for developing an eco-nomic porosity system existed.Therefore, a Type 1 reservoir withthe burial history shown in (A)would be considered far less riskythan the same facies with the burialhistory shown in (B). φ = porosity.

Page 27: Reservoir Quality Prediction in Sand and Carbonates

quartz, authigenic clays (especially chlorite), and cal-cite. Pore-filling iron oxides and other weathering by-products are also commonly associated with ancientsubaerial exposure for both carbonates and clastics.

Exploration risks associated with Type 5 lithofaciesare generally high because their pore systems aredestroyed before hydrocarbon migration can occur.However, risks can be significantly lower depending onthe timing, distribution, and chemical stability (solubil-ity) of the cement phases in question. Therefore, riskassessment necessitates the prediction of two potentialporosity-retaining scenarios, including (1) lateralpinchout of early cements elsewhere, and (2) ancientnear-surface or burial dissolution of cements, or unsta-ble framework grains engulfed within those cements.

Late Burial Cementation (Rock Type 6)

Type 6 lithofacies are originally porous rocks thathave been tightly cemented during late diagenesis by avariety of deep burial cements (Figure 8). Intergranularvolumes in Type 6 lithofacies are usually lower thanthose found in Type 5 lithofacies because of the delayedcementation that accompanies deep burial compaction.For carbonate rocks, the most common burial cementsinclude coarse equant to poikilotopic calcite (both fer-roan and nonferroan), anhydrite, nonplanar dolomite(baroque or saddle), ferroan dolomite, and ankerite.Some of these cements may be hydrothermal in origin,and may also include a variety of accessory mineralssuch as fluorite, galena, pyrite, and sphalerite. In addi-tion to the same burial and hydrothermal cementsfound in carbonate rocks, sandstones may contain bur-ial quartz, feldspar (usually albite), zeolites, and authi-genic clays (kaolinite, illite, smectite).

In general, burial cements have a tendency to be morepervasive, laterally continuous, and chemically stablethan near-surface cements. Therefore, the exploration

risks associated with Type 6 lithofacies may be higher,particularly if the strata in question are thermallymature (dry-gas preservation window or above). Athigh maturity levels, no viable porosity-creating mecha-nisms exist to dissolve these cements (Tobin, 1991).However, risks can be significantly lower, depending onthe timing and distribution of the cement phases inquestion. Some possible porosity-retaining scenariosinclude the following: (1) If the cements are hydrother-mal in origin, they could be laterally restricted to faultzones, bedding contacts, or certain high-permeabilitycarrier beds, and therefore could pinch out laterally intoporous reservoir facies. (2) Ancient near-surface dissolu-tion of burial cements may create secondary porosityelsewhere. (3) Ancient subsurface dissolution of burialcements could also create secondary porosity in Type 6facies, but the probability of an effective dissolutionmechanism is low, particularly if the strata are thermallymature. (4) Productive diagenetic traps (Rittenhouse,1972; Wilson, 1977; Cant, 1986) could be predicted in thesubsurface, especially if it can be demonstrated that oilmigration has occurred prior to or during the initialstages of burial cementation, and the rock samples weretaken from an area that lacked structural or stratigraphicclosure at the time of migration. Burial cements thathave very light δ13C isotope signatures or contain fluo-rescing oil-filled fluid inclusions could represent pastmigration pathways that lead to productive diagenetictraps in higher structural positions in the subsurface.

Hybrid Lithofacies

Nearly all outcrop samples exhibit characteristics oftwo or more of the six rock types described, althoughone characteristic usually dominates. Some unusualhybrids can be found (e.g., a Type 1, tight depositionalrock facies containing fractures that were cemented

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 9

(A) (B)

Figure 4. Photomicrographs illustrating the use of ultraviolet fluorescence to distinguish original depositionalfabric in Type 2 lithofacies. (A) Plane-transmitted light view of a coarse crystalline, nonplanar dolomite lack-ing any obvious depositional texture (Devonian age, Canada). (B) Ultraviolet fluorescence view of the samearea showing a well-defined, grain-supported skeletal fabric containing a variety of open-marine fossils suchas the large punctate brachiopod shell in the lower left. Note also the pressure solution seam between the bra-chiopod and mollusk fragment (arrows). Both photographs were taken at 80×.

Page 28: Reservoir Quality Prediction in Sand and Carbonates

10 Tobin

during recent outcrop exposure), but the most com-monly recognized hybrids are the partially compactedand partially cemented reservoir rocks (Type 4 andType 5/6 combination) and the early and late cementhybrids (Type 5 and 6). For hybrid lithofacies like these,petrographic point-count analyses are essential fordetermining the volumetric contribution of each mech-anism toward total porosity destruction. Porositypotential in the subsurface must be predicted from acombination of the scenarios for each mechanism (com-paction, near-surface cementation, burial cementation).

POROUS ROCKSDecision: Degree of Recent Weathering?

All outcropping strata have been exposed to somedegree of surface weathering. The duration of weath-ering may range from a just a few years to hundreds ofthousands or even millions of years. Surface leachingprocesses can be minimal, or they can create signifi-cant amounts of secondary porosity that are not repre-sentative of the true subsurface conditions that existedprior to recent exposure. This possibility must be con-sidered in risk assessment, particularly with regard toany petrophysical analyses of outcrop materials. Theeffects of surface dissolution can, of course, be mini-mized by avoiding heavily weathered exposures andby using a hammer (or a portable coring device) toobtain the freshest, least altered bedrock below thezone of intense weathering. However, if the efficacyof sample selection is uncertain, several factors shouldbe considered in evaluating the probability of recentdissolution. These include the age of the outcrop (Is ita fresh roadcut? Or a mountain flank, fault escarp-ment, or stream cut exposed for the last 200,000

years?), the prevailing climatic conditions in the area(arid desert outcrops or tropical streamcuts?), outcropproximity to human-induced weathering conditions(e.g., proximity to cultivated farmland with acidicgroundwater runoff), and petrographic evidence ofrecent leaching (Table 5).

If the outcrop sample in question is deemed to be“porous” (i.e., effective porosity and permeability areabove economic requirements), the next question to beasked is “How much surface weathering (and sec-ondary porosity creation) has occurred?” (Figure 1).Three possibilities exist: (1) The outcrop has sustainedminimal weathering, and most of the porosity found isinherited from the subsurface (Type 7 rocks); (2) out-crop weathering is substantial, and most of the poros-ity observed is the result of recent dissolution (Type 8and 9 rocks); or (3) some recent weathering porosity isobserved, but the amount is uncertain (Type 10 rocks).All three possibilities are discussed below.

Recent Weathering Minimal (Rock Type 7)This group includes any reservoir rock whose pore

system has survived intact throughout both burial andrecent outcrop diagenesis (Figure 1). The pore systemof Type 7 rocks represents indigenous porosity inher-ited from the subsurface, and may include not onlyprimary inter- or intragranular pores but also sec-ondary pores that were created by either near-surfaceor subsurface dissolution in the geologic past. It istherefore critical that inherited secondary porosity bedistinguished from secondary porosity created duringrecent outcrop weathering. (Criteria for the recogni-tion of recent weathering-related porosity are dis-cussed in the section on rock Types 8 and 9.

This facies carries the least amount of exploration riskof any of the ten categories described in this chapter.

Figure 5. Example of a Type 3 lithofacies (recent poredestruction) from the Ombilin Basin of Indonesia.This outcrop sample is a sublitharenite sandstonethat has been pervasively cemented by hematite(black opaques) during recent outcrop weathering.Total porosity was reduced from 21% to 4% by recentcements (plane-transmitted light, 125×).

Figure 6. Example of a Type 4 lithofacies (tightlycompacted) from the Chuxiong Basin, China. Thissample is a very tightly compacted Triassic litharen-ite sandstone containing only trace amounts ofmicroporosity. Intergranular volume for this sampleis only 8%, indicating extreme mechanical andchemical compaction (pressure solution).

Page 29: Reservoir Quality Prediction in Sand and Carbonates

Because of the limited weathering involved, samplesfrom this facies are highly suitable for many types ofroutine and special core analyses. Questions that shouldbe addressed as part of risk assessment are: (1) Howmuch porosity and permeability are present? Whatother petrophysical properties can be determined fromthis facies? (2) What was the maximum pre-outcrop bur-ial depth for this facies? (3) How much deeper could thisfacies have been buried before compaction would havedestroyed economic porosity? (4) Is there any reason tosuspect that early overpressuring could exist in the sub-surface that could enhance porosity at depth? (5) Do pet-rographic observations detect any incipient destructivediagenesis that could be more intense at greater burialdepths (or that might be laterally restricted at the samedepth)? Is there any new diagenesis that can be pre-dicted? (6) Is there any potential for further porosityenhancement from ancient near-surface or burial disso-lution? What leaching mechanisms are likely? Whatwould the pore system of this facies look like after disso-lution? What would be the most likely porosity and per-meability? (7) Based on available seismic or well data,how deep are potential traps (prospects) in the basin?Are they deeper than this sample has been buried priorto exposure? What is the probability of finding economicporosity at this depth (Table 6)?

Recent Surface Weathering Dominant(Rock Types 8 and 9)

This group includes any potential reservoir rockwhose pore system is dominated by secondary poros-ity developed during recent outcrop exposure (Figure1). Because of the intense weathering involved, such

samples are not suitable for routine or special coreanalyses. Therefore, any estimates of subsurfaceporosity must be predicted by less direct means, asoutlined below.

The distinction between Type 8 and Type 9 litho-facies is based on the original depositional fabric of therock. Type 8 rocks have depositional fabrics that wereoriginally porous, but have been subsequentlydestroyed by intense cementation or compaction priorto outcrop exposure and leaching. Therefore, theserocks should be considered as equivalents to eitherType 4 (compacted) or Type 5/6 (cemented) lithofacies,depending on the predissolution rock fabric. Petro-graphic identification of secondary pore types andintergranular volume (IGV) can be used to distinguishbetween these two end members. Accordingly, riskassessment should follow the procedures outlined pre-viously for Types 4, 5, and 6 lithofacies, with one excep-tion: the amount of ancient secondary porosity creation,regardless of mechanism, could be similar to that cre-ated during recent outcrop exposure (assuming that thesame rock components have been dissolved, and to thesame extent). Thus, laboratory-measured porosity andpermeability values from weathered outcrop samplescould be representative of subsurface conditions thatmight exist if ancient dissolution has actually occurred.

Type 9 rocks have originally tight depositional fabricsthat have remained tight throughout most of their burialhistory but have been subjected to surface leachingprocesses during recent outcrop exposure. Therefore,these rocks should be considered as equivalents to Type1 rocks (tight depositional facies), and risk assessmentshould follow the precedures outlined for this facies.

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 11

Table 4. Porosity Risk Assessment* for a Type 3 Reservoir Example.†

Question Data Available Answer

Pre-outcrop porosity? Petrographic point count 14% macroporosityEstimated permeability? P vs. K crossplot from analog 50–70 md

in adjacent basinPre-outcrop burial depth? Best-analog compaction curve 2 km

(from Pittman and Larese, 1991, their figure 20)

How much deeper to Best-analog compaction curve (from Probable loss of porosity to 10%economic porosity basement? Pittman and Larese, 1991, their by 2.5 km

figure 20); assumes a 10% economic porosity cutoff

Overpressuring or pressure No incipient pressure solution noted; Low probability of either solution likely? framework composition not conducive

to pressure solution; no overpressures observed in adjacent basin

Additional secondary Petrographic description Potential dissolution of unstable porosity likely? lithics would add another 8%

porosity

*Risk assessment:• economic porosity basement (10%) likely to be encountered at 2.5 km if no secondary dissolution• economic porosity basement (10%) likely to be encountered deeper (>4.5 km) if secondary porosity is present• minimum porosity likely at 2.5 km = 10%; maximum = 18%.

†Iron oxide cemented lithic sandstone.

Page 30: Reservoir Quality Prediction in Sand and Carbonates

12 Tobin

Degree of Recent Weathering Uncertain(Rock Type 10)

This category includes any reservoir lithologywhose pore system contains appreciable amounts ofsecondary porosity of uncertain origin. Not surpris-ingly, most porous outcrop samples fall into this cate-gory, primarily because much of the physical evidencefor recent dissolution is equivocal (Table 5), has beenmasked by a variety of diagenetic by-products, or ismissing altogether. For these samples, the degree of riskassociated with reservoir porosity remains uncertain.

CASE HISTORIESChuxiong Basin, Yunan Province, China

In this study, sandstone outcrops of Upper Triassicage were sampled and petrographically evaluated byTobin and Nelis (1990) in an effort to characterizepotential reservoir quality in two of Amoco’s prospects.Both structures are interpreted as having been upliftedfrom a maximum burial depth of ~5 km to their presentdepth of ~3 km (based on available seismic data, sedi-ment thickness estimates, and basin modeling). Giventhe thickness and areal extent of sandstone facies in thisarea, the average minimum porosity required for aneconomic gas play would be 12%.

Although a few examples of Type 1 (tight deposi-tional facies) and Type 10 (abundant secondary poros-ity of uncertain origin) lithofacies are present, the vastmajority of the outcrop samples collected are classifiedas Type 4 (nonporous, tightly compacted) and Type 8(tightly compacted, but porous and weathered) rocks(Figure 1). Most of these samples are immaturelitharenites, feldspathic litharenites, or lithic arkosesthat have suffered extreme primary porosity loss fromintense mechanical and chemical compaction. Inter-granular volume for these facies ranges from8% to 29% (mostly 8%–12%), and intergranularcements are minimal, ranging from 3% to 5%.

The Type 8 lithofacies examined, although porous,exhibit unmistakable evidence for intense recent out-crop leaching, including abundant iron oxide staining,soil formation above outcrops, abundant iron oxidecoatings in secondary pores, iron oxide rims andcleavage traces floating in secondary pores, and theubiquitous occurrence of secondary pores engulfedwithin highly compacted rock fabrics. Type 8 rocksamples are texturally and mineralogically equivalentto the Type 4 (tightly compacted) lithofacies from thesame area, but are only exposed in outcrops that aredownstream from cultivated farmland. It is believedthat the higher acidic groundwater runoff associatedwith these types of exposures is responsible for thepreferential dissolution observed. Therefore, theexploration risks associated with this facies are consid-ered to be the same as that of Type 4 rocks.

Figure 7. Example of a Type 5 lithofacies (early near-surface cemented). This Jurassic outcrop samplefrom Somalia is an ooid lime grainstone that wastightly cemented during early diagenesis by bladedand equant calcite (plane-transmitted light, 40×).

Figure 8. Two examples of Type 6 lithofacies (burialcemented) from Trinidad. (A) A skeletal lime grain-stone containing red algae (R), forams (F), and mol-lusks (M) is tightly cemented by poikilotopic ferroancalcite of burial origin. (B) A quartzarenite sandstoneis tightly cemented by nonplanar ferroan dolomite ofburial origin (plane-transmitted light, 80×).

(A)

(B)

Page 31: Reservoir Quality Prediction in Sand and Carbonates

By far, the most abundant rock type observed is theType 4 variety. Porosities observed in this facies are con-siderably lower than the 12% required for an economicplay (0–5%, mostly <3%). Therefore, this group is con-sidered to be a very high-risk exploration target, unless amechanism can be predicted for primary porositypreservation or secondary dissolution elsewhere.

Preservation of Primary PorosityEarly overpressuring can retard the rate of porosity

loss from compaction, but shallow overpressures arenot known to occur in this area. Similarly, early grain-coating rim cements can also retard compaction, butonly minor amounts (0–6%, mostly <1%) of early quartzovergrowth cement are present in the outcrop samplesexamined. Furthermore, the high lithic content wouldprobably limit the heterogeneous nucleation of quartzin these sandstones elsewhere, thereby reducing theporosity-preserving effectiveness of cementation. Alter-natively, primary porosity could be more extensive inprospects that are shallower than pre-outcrop burialdepth. However, vitrinite reflectance data indicate thatpre-outcrop burial depth for the samples examined was~4 km (based on a paleogeothermal gradient of 0.2%Ro/km and an assumed surface intercept of 0.2% Ro).Using an unrealistically optimistic linear compactionmodel, these samples are interpreted to have reachedtheir economic porosity basement of 12% at a depth of3 km, ~2 km less than the maximum burial depth (5 km)sustained by the two Amoco prospects (Figure 9).Experimental compaction studies, however, indicatethat a linear compaction model is unrealistic; the effectof cementation and both mechanical and chemical com-paction would be to reduce porosity to its economicbasement at much shallower depths (Pittman andLarese, 1991). Thus, economically viable primary poros-ity preservation in this area is highly unlikely.

Creation of Secondary PorosityAncient dissolution of some of the unstable feldspars

and lithic grains in these strata could yield economicporosity elsewhere. Two burial dissolution mechanismsare possible: dissolution by undersaturated water derivedfrom shale compaction, or organic acid dissolution. Theformer is considered unlikely because the strata in ques-tion have been buried to a depth of ~4 km with no obvi-ous signs of grain dissolution. If compaction waterleaching had occurred, it would have taken place at con-siderably shallower depths. The latter mechanism(organic acids) is also considered unlikely for this area,because the undissolved sandstones examined are inter-calated with organic-rich shales that have maturedenough to have generated oil (~1.0% Ro), a level of matu-rity well past what is required for organic acids to formand migrate. The absence of organic acid dissolution mayalso be, in part, the result of two other factors: (1) theinterbedded source rocks are dominated by gas-proneType III kerogen, a potentially poor source of liquids(including organic acids); and (2) any acids or other typesof undersaturated pore fluids that might have reached thesandstones in question would likely have been somewhatineffective at creating secondary porosity because of thelack of an open, permeable pore system (destroyed duringearly burial by compaction). The only realistic mechanismfor ancient dissolution would be meteoric (near-surface)leaching associated with paleoexposure surfaces. Ancientnear-surface leaching is more likely to create significantsecondary porosity because of the higher rock surfacearea exposed, exposure-related fracturing and pressureunloading, and the higher fluid flow rates involved.Therefore, porous sandstone reservoirs might exist belowpaleoexposure surfaces (unconformities). Accordingly,unconformity-related prospects may be less risky than thestructural prospects previously identified.

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 13

Table 5. Petrologic Criteria for Distinguishing Recent Outcrop Dissolution.

Observational Scale Level of Confidence Description

Megascopic Suggestive Abundant iron oxide staining on outcropsMegascopic Suggestive Soil or caliche formation on outcropsMegascopic Diagnostic Recent karstic landforms and associated secondary porosityMegascopic Diagnostic Soft, porous, weathered rims on outcrop surfaces with hard,

tight rock beneathMacroscopic Diagnostic Gradational dissolution rims on hand specimensMicroscopic Suggestive Abundant iron oxide coatings associated with secondary

poresMicroscopic Diagnostic Iron oxide rims or cleavage or grain fracture traces "floating"

in secondary poresMicroscopic Suggestive Late, postcompaction secondary pores in an otherwise

tightly compacted rockMicroscopic Suggestive Secondary pores that postdate deep burial cements,

fractures, or stylolitesMicroscopic Diagnostic Secondary pores that postdate recent surface to near-surface

cementsMicroscopic Diagnostic Secondary pores that postdate oil entrapment by-products

(e.g., bitumen)

Page 32: Reservoir Quality Prediction in Sand and Carbonates

14 Tobin

Disposition of Prospect

Because of the high risks associated with reservoirquality in the subsurface, the two Amoco prospectsunder evaluation were not drilled. Subsequent wellsdrilled in this area by the Chinese have not penetratedthe Triassic sandstone.

West-Central Myanmar

In an effort to assess the degree of reservoir riskprior to drilling, Murphy et al. (1991) described sand-stone outcrop samples from the Paunggyi Formation(Paleocene to early Eocene in age) in the ChindwinBasin of Myanmar (Burma). The depth of the reservoirat the drilling prospect (the Yenan structure in BlockB) was estimated to be ~6500 ft (2 km). Given the thick-ness and areal extent of sandstone facies in this area,the average porosity required for an economic oil playwould have been ~15%.

Approximately two-thirds of the outcrop samplesdescribed in this study are immature litharenites andfeldspathic litharenites containing minor amounts ofporosity (mostly <3%). These samples include tightlycompacted Type 4 lithofacies and Type 4/6 hybridswhose pore systems were destroyed by a combina-tion of compaction and burial cementation (mostlycalcite, dolomite and siderite, and minor quartz). Theremaining samples are more mature quartzarenites,

sublitharenites, and subarkoses. These are classifiedas Type 10 (recent weathering uncertain) rocks (forthose containing more than 15% porosity), or asType 5 (early near-surface cemented), Type 6 (lateburial cemented), or Type 4/6 (compacted/burialcemented hybrid) rocks (for less porous examples).Intergranular volume is slightly lower for the imma-ture sample group (mostly <25%) and higher for themore mature group (mostly >25%). Intergranularcements range from 2% to 32%.

Exploration risks are considered to be high for theimmature, highly compacted sandstone facies (Type 4and Type 4/6) in the Paunggyi Formation, unless pri-mary porosity preservation or secondary dissolutioncan be predicted elsewhere. Early overpressuring is apotential mechanism for retarding the rate of porosityloss from compaction; overpressures are known toexist in the basin from previous drilling reports. How-ever, the timing and both lateral and vertical extentsof overpressure are uncertain. Similarly, early grain-coating rim cements can also retard compaction, butonly minor amounts of early cement (quartz andsome calcite) are present in the outcrop samplesexamined, and they do not appear to have signifi-cantly reduced the amount of porosity loss from com-paction (based on IGV data). Alternatively, porositycould be higher for this facies if the drilling prospectis significantly shallower than pre-outcrop burial

Table 6. Porosity Risk Assessment* for a Type 7 Reservoir Example.

Question Data Available Answer

Measured porosity Petrographic point count + routine 14% macroporosity, 110 md and permeability? core analysis permeability

Pre-outcrop burial depth? Best-analog compaction curve 1.5 km (from Pittman and Larese, 1991,their figure 20)

How much deeper to Best-analog compaction curve Probable loss of porosity to 10%economic porosity basement? (from Pittman and Larese, by 2.5 km; 8% by 3 km

their figure 20); assumes a 10% economic porosity cutoff

Overpressuring or pressure No incipient pressure solution noted; Low probability of either solution likely? framework composition not

conducive to pressure solution; no overpressures observed in adjacent basin

Any destructive diagenesis Petrographic description No incipient burial cements notedlikely?

Additional secondary porosity Petrographic description Potential dissolution of unstable likely? lithics would add another 6%

porosityHow deep are prospects Seismic data only Structural traps at 3 km

in the basin?

*Risk assessment:• economic porosity basement (10%) likely to be encountered at 2.5 km if no secondary dissolution• prospects at 3 km, likely primary porosity remaining = 8%• potential for additional secondary porosity of 6%• minimum porosity likely at 3 km = 8%; maximum = 14%.

Page 33: Reservoir Quality Prediction in Sand and Carbonates

depth. However, vitrinite reflectance data from thisinterval (0.5% Ro) indicate that pre-outcrop burialdepth (7500 ft; 2.3 km) was only about 1000 ft (0.3 km)deeper than the drilling prospect (based on a paleo-geothermal gradient of 0.04% Ro/1000 ft and anassumed surface intercept of 0.2% Ro). Furthermore,experimental compaction studies by Pittman andLarese (1991, their figure 21) indicate that the 15%porosity basement for immature lithic sands like thesewould be considerably shallower than the drillingprospect [at ~2700 ft (0.8 km)] (Figure 10). This estimatedoes not include the additional risk associated withconcomitant burial cementation observed in some ofthe outcrop samples. Dissolution of these cements(and/or unstable framework grains) is not consideredto be a realistic mechanism for creating additionalporosity because of the poor plumbing system that ischaracteristic of these highly compacted rocks.

Exploration risks are lower for the more maturesandstone facies (Types 5, 6, and 10). These sandstoneshave a less ductile framework composition, and conse-quently less compaction, but porosity values are stillmostly below the 15% economic limit because of theeffects of burial cementation. However, dissolution ofcements and/or framework grains is more likely to be

an effective mechanism for creating additional sec-ondary porosity for these samples because of theirmore permeable plumbing system (for circulatingundersaturated pore fluids). Indeed, some of the sand-stones in this group (Type 10 facies) contain an appre-ciable amount of secondary porosity, but its origin(ancient or recent?) is uncertain. Either way, the combi-nation of primary and secondary porosity for thisfacies could be in the 15%–20% range. Ancient subsur-face dissolution and near-surface leaching associatedwith paleoexposure are viable mechanisms for creatingsecondary porosity for these facies.

In order to increase the odds of drilling success inthis basin, Murphy et al. (1991) recommended that (1)a regional provenance study be conducted to map thelocalities of the more mature sandstone facies, and (2)seismic data be used to identify potential subaerialexposure surfaces. These criteria were intended tomatch the best sandstone compositions with the high-est probability of ancient near-surface dissolution.

Disposition of ProspectBecause of the high risks associated with reservoir

presence and quality in the subsurface, the Yenanprospect was not drilled. However, a second prospect

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 15

4 03 53 02 52 01 51 050

6 . 0

5 . 5

5 . 0

4 . 5

4 . 0

3 . 5

3 . 0

2 . 5

2 . 0

1 . 5

1 . 0

0 . 5

0 . 0

Porosity (%)D

epth

(k

m)

12 % minimum porosity projected at3 km based on linear model

linear c

ompaction m

odel for o

utcrop sa

mplePittman and Larese (1991) model(corrected for cementation andpressure solution )

maximum burial depth sustained by Amoco prospects = 5 km(porosity would be 0 % if compaction followed linear model; < 10 % if compaction followed Pittman and Larese, 1991, model)

economic porosity(above 12 % cutoff)

uneconomic porosity(below 12 % cutoff)

12 % minimum porosityprojected at 1.5 km fromPittman and Larese(1991) model

Figure 9. Prediction of pre-served primary porosity fromcompaction modeling in theChuxiong Basin (China)example. The linear model(an unrealistically optimistictool) predicts loss of porosityto the economic minimum(12%) at 3 km, which is thepresent-day depth of thedrilling prospects but is2 km less than maximum bur-ial depth (5 km). The Pittmanand Larese (1991) model (cor-rected for cementation andpressure solution) suggeststhat the economic porosityminimum would be encoun-tered at much shallowerdepths (1.5 km). Therefore,economic porosity should notbe expected in the drillingprospects.

Page 34: Reservoir Quality Prediction in Sand and Carbonates

16 Tobin

was drilled in the same block in early 1992 (Amoco #1Uyu). This prospect encountered low-porosity lithicsandstone of slightly younger age (Miocene) whoseframework composition is analogous to the Paunggyisandstones described by Murphy et al. (1991). The poresystem of this sand was greatly reduced below 2 km to<8% by a combination of mechanical compaction andcementation (mainly carbonates, zeolites, and authi-genic clays). This mode of pore destruction closelymatches predictions made by Murphy et al. (1991) inaddition to three other independent Amoco studies ofoutcrop samples in Block B (Taylor et al., 1993).

Central Taurids, Turkey

In this study, carbonate outcrops of Jurassic age werepetrographically studied by Tobin (1992) as part of anearly risk evaluation of the central Taurids in south-western Turkey. Two potential reservoirs weredescribed: dolomites from the Jurassic Hendos Forma-tion and limestones from the Jurassic Pisarcukuru For-mation. The Hendos samples are medium to coarsecrystalline, planar dolomites lacking any evidence oftheir original depositional fabric. Porosity for this groupranges from 2% to 6%, with permeabilities of <0.2 md.Therefore, these dolomites were initially considered

Type 2 rocks (uncertain depositional facies), with anassociated high degree of exploration risk. Later petro-graphic data, however, suggested that this facies couldhave higher porosity and permeability values under thefollowing conditions: (1) Incipient fracturing of this brit-tle lithology was observed in thin section. More intensefracturing elsewhere could increase the porosity by afew percentage points, and would greatly increase per-meability. (2) Incipient paleokarst (vuggy and skel-moldic secondary porosity) was observed in allsamples, suggesting the potential for higher porositieswherever karstic dissolution was more pervasive. (3)The rock matrix consists of planar dolomite crystals ofnear-surface to shallow burial origin and late planar tononplanar dolomite of suspected burial origin. If theburial dolomite postdates hydrocarbon migration, thisfacies could contain up to 12% porosity (based on thin-section point-count data) in structures that existed atthe time of migration because of the cement retardingeffect of hydrocarbons (Wilson, 1977).

The Pisarcukuru Formation samples are medium-grained, well-sorted ooid-skeletal lime grainstonescontaining <1% porosity and 0.01 md of permeability.Porosity loss is the result of both compaction (63%;both mechanical and chemical) and early, near-surface

15 % minimum porosityprojected at 0.8 km fromPittman and Larese (1991)

4 03 53 02 52 01 51 050

6 . 0

5 . 5

5 . 0

4 . 5

4 . 0

3 . 5

3 . 0

2 . 5

2 . 0

1 . 5

1 . 0

0 . 5

0 . 0

Porosity (%)D

epth

(k

m)

linear compaction model for o

utcrop samplePittman and Larese (1991) model

economic porosity(above 15 % cutoff)

uneconomic porosity(below 15 % cutoff)

Amocoprospect(2 km)

15 % minimum porosityprojected at 1.9 km fromlinear model

Figure 10. Prediction of pre-served primary porosity fromcompaction modeling in theChindwin Basin (Myanmar)example. The linear model(an unrealistically optimistictool) predicts loss of porosityto the economic minimum(15%) at 1.9 km, slightly lessthan the present-day depth ofthe drilling prospect (2 km).The Pittman and Larese (1991)model suggests that the economic porosity minimumwould be encountered atmuch shallower depths (0.8 km). Therefore, economicporosity should not be expected in the drillingprospect.

Page 35: Reservoir Quality Prediction in Sand and Carbonates

cementation (37%; including isopachous micritic,isopachous bladed, syntaxial, and minor equantcements). Thus, these samples are considered Type 4(compacted) and Type 5 (early near-surface cemented)hybrid lithofacies. The exploration risks associatedwith reservoir quality were initially regarded as highbecause of low porosity and permeability. Petro-graphic data, however, suggest that exploration risk ismoderate for this lithofacies, because of the potentialfor the following: (1) lateral cement pinchout couldresult in up to 15% intergranular porosity, and (2) thedissolution of chemically unstable grains (includingooids, some foraminifera, and mollusk fragments)could have contributed additional porosity elsewhere.

Disposition of Play

Because of a variety of technical risk factors, thisplay was discontinued by Amoco, and no drillingprospects were recommended.

FUTURE RESEARCH

Predicting subsurface porosity and permeabilityfrom outcrop materials is risky business. To be success-ful, both ancient burial history and associated diagene-sis as well as recent diagenesis and associated porositymodification must be accurately determined from pet-rographic or geochemical clues preserved in the rocks.These preserved signposts of complete diagenesis andporosity evolution are in part straightforward and inpart extremely subtle. The weak links in this evaluationinvolve recognizing ancient and recent diagenesis forwhich petrographic criteria are limited or equivocal,such as petrographic criteria for recognizing and quan-tifying recent leached porosity, and criteria for recog-nizing recent pore-filling cements. Another weak link isthe estimation of potential permeability in cementedreservoir rocks (Type 5 and Type 6 lithofacies) whosecements are predicted to be absent because of lateralpinchout or ancient dissolution. These areas of investi-gation are considered fertile ground for future research.

CONCLUSIONS

Outcrop observations can greatly assist in reservoirrisk assessment, particularly in frontier basins wheresubsurface data are sparse. Outcrop exposures pro-vide a three-dimensional view of sedimentary facies inaddition to unlimited rock sample availability for lab-oratory analyses. The systematic, decision-treemethodology outlined in this chapter can greatlyenhance the efficiency and completeness of outcrop-based reservoir prediction studies.

Every potential reservoir rock exposed in outcrophas a pore system that is the end product of its originaldepositional facies and subsequent diagenetic history,including both pre-exposure diagenesis and recentweathering effects. Therefore, reliable risk assessmentmust consider depositional and diagenetic history,including both ancient near-surface and subsurfacediagenesis and recent weathering.

The discovery of tight reservoir facies in surfaceoutcrop exposures does not necessarily mean that ahigh degree of risk should be assigned to subsurfaceporosity preservation. Tight facies may be assigned alow degree of risk whenever the following diageneticconditions can be predicted: (1) recent destructive dia-genesis of originally porous facies, (2) natural subsur-face fracturing, (3) dolomitization of low-porositylimestone, (4) drilling prospect depths that are suffi-ciently shallower than the outcrop analog to preserveporosity, (5) early overpressuring or early thin rimcementation, (6) ancient secondary dissolution, (7)ancient karsting, (8) lateral or vertical cement pinch-out in the subsurface, or (9) petroleum inhibition ofcementation or other destructive diagenesis.

The discovery of porous reservoir facies at the sur-face does not necessarily guarantee that the samerocks will be porous in the subsurface. Recent surfaceleaching or fracturing can create secondary porositythat is not likely to exist in the same formation in thesubsurface.

ACKNOWLEDGMENTS

The author thanks Dick Larese, Ron Nelson, andPaul Wagner for constructive reviews of an earlier ver-sion of this manuscript. Thanks are also extended toAAPG reviewers Andrew Leonard, Pascual Marquez,and Jon Gluyas. Many of the concepts presented inthis paper are the by-products of fruitful conversationswith fellow colleagues, in particular Dick Larese, IoneTaylor, Mary Nelis, Tim Murphy, and Christine Skir-ius. I am indebted to you all. Thanks also to Amoco forpermission to publish this work.

REFERENCES CITED

Bloch, S., 1991, Empirical prediction of porosity andpermeability in sandstones: AAPG Bulletin, v. 75, p. 1145–1160.

Bloch, S., and K.P. Helmold, 1995, Approaches to pre-dicting reservoir quality in sandstones: AAPG Bul-letin, v. 79, p. 97–115.

Byrnes, A.P., 1994, Empirical methods of reservoirquality prediction, in M.D. Wilson, ed., Reservoirquality assessment and prediction in clastic rocks:SEPM Short Course 30, p. 9–21.

Cant, D.J., 1986, Diagenetic traps in sandstones: AAPGBulletin, v. 70, p. 155–160.

Dravis, J.J., and D.A. Yurewicz, 1985, Enhanced car-bonate petrography using fluorescence microscopy:Journal of Sedimentary Petrology, v. 55, p. 795–804.

Dunnington, H.V., 1967, Aspects of diagenesis and shapechange in stylolitic limestone reservoirs: Elsvier Pub-lishing Co. Ltd. 7th World Petroleum Congress Pro-ceedings, April 2–8, Mexico, v. 2, p. 339–352.

Feazel, C.T., and R.A. Schatzinger, 1985, Prevention ofcarbonate cementation in petroleum reservoirs, inN. Schneidermann and P.M. Harris, eds., Carbonatecements: SEPM Special Publication 36, p. 97–106.

Porosity Prediction in Frontier Basins: A Systematic Approach to Estimating Subsurface Reservoir Quality 17

Page 36: Reservoir Quality Prediction in Sand and Carbonates

18 Tobin

Folk, R.L., 1987, Detection of organic matter in thinsections of carbonate rocks using a white card: Sed-imentary Geology, v. 54, p. 193–200.

Goldstein, R.H., 1988, Cement stratigraphy of Pennsyl-vanian Holder Formation, Sacramento Mountains,New Mexico: AAPG Bulletin, v. 72, p. 425–438.

Kerans, C., 1988, Karst-controlled reservoir hetero-geneity in Ellenburger Group carbonates of WestTexas: AAPG Bulletin, v. 72, p. 1160–1183.

Lucia, F.J., and R.P. Major, 1994, Porosity evolutionthrough hypersaline reflux dolomitization, in B.H.Purser, M.E. Tucker, and D.H. Zenger, eds.,Dolomites: a volume in honor of Dolomieu: Inter-national Association of Sedimentologists SpecialPublication 21, p. 325–341.

Meshri, I.D., 1989, On prediction of reservoir qualitythrough chemical modeling (abs.): AAPG Bulletin,v. 73, p. 390–391.

Meshri, I.D., and P.J. Ortoleva, 1990, Prediction ofreservoir quality through chemical modeling:AAPG Memoir 49, 175 p.

Moore, C.H., 1989, Carbonate diagenesis and porosity:Amsterdam, Elsevier, Developments in Sedimen-tology 46, 338 p.

Moore, C.H., and Y. Druckman, 1981, Burial diagenesisand porosity evolution, Upper Jurassic Smackover,Arkansas and Louisiana: AAPG Bulletin, v. 65, p. 597–628.

Murphy, T.B., R.C. Tobin, and W.W. Dorsey, 1991,Reservoir risk assessment of Paleocene–LowerEocene outcrop samples from west-central Myan-mar: unpublished Amoco report, 51 p.

Nelson, R.A., 1985, Geological analysis of naturallyfractured reservoirs: Houston, Gulf PublishingCompany, 320 p.

Ottmann, R.D., P.L. Keyes, and M.A. Ziegler, 1976, JayField, Florida—a Jurassic stratigraphic trap, in J.Braunstein, ed., North American oil and gas fields:AAPG Memoir 24, p. 276–286.

Pittman, E.D., and R.E. Larese, 1991, Compaction oflithic sands: experimental results and applications:AAPG Bulletin, v. 75, p. 1279–1299.

Purser, B.H., 1978, Early diagenesis and the preserva-tion of porosity in Jurassic limestones: Journal ofPetroleum Geology, v. 1, p. 83–94.

Purser, B.H., A. Brown, and D.M. Aissaoui, 1994,Nature, origins and evolution of porosity indolomites, in B.H. Purser, M.E. Tucker, and D.H.Zenger, eds., Dolomites: a volume in honor ofDolomieu: International Association of Sedimentol-ogists Special Publication 21, p. 283–308.

Rittenhouse, G., 1972, Stratigraphic-trap classification,in R.E. King, ed., Stratigraphic oil and gas fields—

classification, exploration methods and case histo-ries: AAPG Memoir 16, p. 14–28.

Saller, A.H., D.A. Budd, and P.M. Harris, 1994, Uncon-formities and porosity development in carbonatestrata: ideas from a Hedberg Conference: AAPGBulletin, v. 78, p. 857–872.

Scherer, M., 1987, Parameters influencing porosity insandstones: a model for sandstone porosity predic-tion: AAPG Bulletin, v. 71, p. 485–491.

Scholle, P.A., L. Stemmerik, and D.S. Ulmer, 1991, Dia-genetic history and hydrocarbon potential of UpperPermian carbonate buildups, Wegener Halvø area,Jameson Land Basin, East Greenland: AAPG Bul-letin, v. 75, p. 701–725.

Surdam, R.C., and L.J. Crossey, 1987, Integrated diage-netic modeling: a process-oriented approach forclastic systems: Annual Review of Earth and Plane-tary Science, v. 15, p. 141–170.

Taylor, I.L., L.E. McRae, and G.O. Smith, 1993, Sedi-mentology and diagenesis of Tertiary sandstonesfrom the Chindwin Basin, Myanmar (Burma): a casestudy for predicting reservoir quality from outcrop(abs.): AAPG Annual Convention Program, April25–28, New Orleans, p. 188–189.

Tobin, R.C., 1985, Reservoir development in the Ellen-burger Group of West Texas: a diagenetic Jambal-aya (abs.): AAPG Bulletin, v. 69, p. 312.

Tobin, R.C., 1991, Pore system evolution vs. paleotem-perature in carbonate rocks; a predictable relation-ship? (abs.): Organic Geochemistry, v. 17, p. 271.

Tobin, R.C., 1992, Petrography and core plug analysesof outcrop samples, central Taurids, southwesternTurkey: unpublished Amoco report, 26 p.

Tobin, R.C., and M.K. Nelis, 1990, Prediction of sub-surface reservoir quality from outcrop samples col-lected in the Chuxiong Basin, Yunan Province,southern China: unpublished Amoco report, 34 p.

Wagner, P.D., and R.K. Matthews, 1982, Porositypreservation in the upper Smackover (Jurassic) car-bonate grainstone, Walker Creek field, Arkansas:response of paleophreatic lenses to burial processes:Journal of Sedimentary Petrology, v. 52, p. 3–18.

Weyl, P.K., 1959, Pressure-solution and the force ofcrystallization: phenomenological theory: Journalof Geophysical Research, v. 64, p. 2001–2025.

Weyl, P.K., 1960, Porosity through dolomitization:conservation-of-mass requirements: Journal of Sed-imentary Petrology, v. 30, p. 85–90.

Wilson, H.H., 1977, “Frozen-in” hydrocarbon accumu-lations or diagenetic traps—exploration targets:AAPG Bulletin, v. 61, p. 483–491.

Wood, J.R., 1994, Geochemical models, in M.D. Wilson,ed., Reservoir quality assessment and prediction inclastic rocks: SEPM Short Course 30, p. 23–40.

Page 37: Reservoir Quality Prediction in Sand and Carbonates

19

Chapter 2

Prediction of Porosity in Compacted SandsJon Gluyas1

BP Exploration de Venezuela SACaracas, Venezuela

Christopher A. CadeBP Norge UA

Stavanger, Norway

ABSTRACT

We present a new porosity–depth relationship for clean, rigid grain(quartz, feldspar) sands under hydrostatic burial. This allows the predictionof porosity in uncemented sandstones to an accuracy of ±2.5 porosity unitsat 95% confidence levels. The relationship was derived using experimentaldata from laboratory compaction experiments and field data for burieduncemented sandstones from around the world. The equation is:

where porosity (φ) is in percentages and depth (z) is in meters.By scaling this relationship in terms of effective stress rather than depth, it

can be used to provide an equally accurate prediction of porosity for uncement-ed sands in overpressured settings. This is done using the following equation:

where z’ = effective burial depth (in meters); z = burial depth (in meters);ρr = density of rock (in Kgm–3 [kilograms per cubic meter]) = typically 2650;ρw = density of water (Kgm–3) = typically 1050; g = gravity (in ms–2 [metersper second squared]) = 9.8; φ∑= average porosity of overburden = typically0.2; and u = overpressure (in MPa [megapascals]).

We propose that there is considerable value in a “compaction only” porosity–depth relationship. A compaction-only trend allows the accurate prediction ofporosity in uncemented sandstones, and gives a maximum porosity baseline towhich cement volumes, and resultant cemented sandstone porosities, can becompared. If both cemented and uncemented sandstone data are included toproduce a “porosity loss–depth” relationship, the resultant scatter (typically ±5porosity units for a given depth) in the relationship limits its usefulness.

z zugr w

' –– –

= ( ) ( )

ρ ρ 1 φΣ

φ =+ ×

50

102 4 5 10

3

4exp–

.

–z

z

Gluyas, J., and C.A. Cade, 1997, Prediction of porosityin compacted sands, in J.A. Kupecz, J. Gluyas, andS. Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69, p. 19–28.

1 Present affiliation: Monument Oil and Gas, London, United Kingdom

Page 38: Reservoir Quality Prediction in Sand and Carbonates

20 Gluyas and Cade

INTRODUCTION

When exploring for oil and gas reservoirs, the pre-diction of porosity before drilling is an important partof the decision-making process. Higher porosity usu-ally gives higher in-place and reserve volumes, andhigher production flow rates (because permeability istypically proportional to porosity). The prediction ofporosity can be made using several different methods:

• Using global, regional, or local porosity–depth curves.This is probably the most commonly used method.For sandstones covering an appropriate depthrange, porosity is plotted against depth, andregression is used to establish a best-fit line orcurve. This line or curve, or the equation thatdescribes it, can then be used to predict porosity forthe undrilled prospect. Examples of such curvestake logarithmic (Athy, 1930; Weller, 1959), powerfunction (Baldwin and Butler, 1985), or linear (Sel-ley, 1978) forms, and they can give accurate poros-ity predictions, particularly when they describeporosity variation with depth for one sandstonetype with a consistent burial and diagenetic charac-ter. When the curve has wider scope (for example,if it is based on data from sandstones of differingage, mineralogy, diagenetic history, or geographiclocation), the resultant scatter in the data will usu-ally increase predictive uncertainty.

• Using a wider group of porosity-controlling variablesto predict porosity. The porosity–depth curve usesdepth as the only control on porosity. Scherer(1987) assembled a diverse group of sandstonesfrom around the world for which there were dataon a range of mineralogical and textural parame-ters, as well as age and depth. Using multipleregression, a predictive equation for porosity wasestablished with a wide group of input variables.Porosity was principally correlated with burialdepth, but there were other important variablesthat determined the degree to which porosity wasreduced for each depth increment. For example, asandstone with numerous ductile grains losesmore porosity, for an equivalent depth of burial,than does a pure quartz sandstone, all other con-ditions being equal. The multiple-regressionmethod of Scherer (1987) attempts to account for awider group of controls, and produces well-constrained relationships between porosity and arange of rock properties. However, the methodhas serious limitations. For example, one of theprimary porosity determinants in Scherer’s data

set is sorting, and in many cases it will be difficultor impossible to predict sorting for an undrilledsandstone with any confidence.

In order to predict porosity from a specific rockproperty, prior to drilling, the particular rock propertymust be known or predictable. Depth to a prospect isusually well constrained, so a porosity–depth curve iseasily applied. However, unless the porosity–depthcurve is based on local data, the prediction uncertaintywill commonly be too large to be useful.

We present an approach to porosity predictionbased on the compaction process and parameters thatare usually predictable: depth and pressure. Theresults of high-pressure laboratory compaction tests onquartzose sands are combined with porosity data froma varied data set of buried and uncemented sands toproduce porosity–depth and porosity–effective stressrelationships for the compaction process.

A third parameter, detrital mineralogy, may also bepredictable in many cases, but in this chapter onlyquartzose (and quartzo-feldspathic) sands are consid-ered. A similar approach applied to lithic sandstonesis the subject of a separate paper in preparation.

This approach focuses on compaction only and istherefore applicable, on its own, to uncemented sands.Most published porosity prediction relationships(Athy, 1930; Selley, 1978; Baldwin and Butler, 1985),consider total porosity loss with burial, which includesboth compaction and cementation; this in partaccounts for the wide range of porosity, at any givendepth, in their data sets. We propose that there is valuein separating compaction and cementation effects,partly to narrow the range of predicted porosity at agiven depth, but also because uncemented sandstonesare a frequent exploration target.

POROSITY LOSS IN SANDSTONES

Recently deposited sands are usually highlyporous, often >40% (Pettijohn, 1975). Buried sandsand sandstones have lower porosities (Table 1).Porosity is reduced by two distinct and commonlyindependent processes: compaction and cementation.The difference between these two processes is mosteasily considered in terms of pore volume and bulkrock volume change. Compaction involves the reduc-tion in pore space associated with shortening of thesand column under burial loading (reduction in bothpore volume and bulk rock volume). Cementation, incontrast, involves a reduction in pore space without

Prior to drilling, the new relationships may be used either to predict theporosity of sands that are known to be uncemented or to place an upperlimit on the porosity estimated for sandstones either known or suspected tobe cemented.

Page 39: Reservoir Quality Prediction in Sand and Carbonates

any reduction of bulk rock volume. Pore space isfilled, partly or completely, by newly precipitatedsolid material.

In general, compaction is the dominant porosity-reduction process during early and shallow burial.Exceptions to this include cemented beach sandstones,calcrete paleosol sandstones, evaporitic sabkha sand-stones, and sandstones that are exposed at a marinesurface during a break in sedimentation. Such exam-ples are not uncommon, but in terms of the sandstonecomponent in the sediment column, they are rarelyvolumetrically significant.

As depth of burial and/or age increases, the relativeimportance of cementation tends to increase. Thereare, however, numerous exceptions to this pattern. Forexample, in the Gulf of Mexico (Table 1), uncementedsands are recorded at depths >3800 m. Also, there aremany instances that demonstrate that the degree ofcementation can vary widely even within the sameformation, depth range, and geographical area. Forexample, in the Central Graben of the North Sea, bothmoderately and almost totally quartz-cemented sand-stones of Upper Jurassic age occur at the same depth(~4000 m), but are areally separated by only a fewhundred meters (Gluyas, this volume; Ramm et al.,this volume). This variability produces much of thescatter on published porosity–depth plots. This scattermeans that these relationships cannot give predictiveaccuracy to better than ±5 porosity units. We suggestthat much of the spread in the publishedporosity–depth curves is derived from the mixing ofcompaction, cementation, and overpressure effects.

There are two reasons for considering compactionon its own. First, uncemented buried sandstones arenot uncommon, and any predictive relationship forporosity in such lithologies should be based on datathat exclude the impact of cementation. Second, wherecemented sandstones are expected, a compaction-onlytrend gives a maximum porosity for a given depth.Deviations to values lower than this may be estimatedfrom predictions of likely authigenic mineral volumes.Such predictions may come from diagenetic modeling,stratigraphic context, regional data, and other meth-ods. Porosity loss due to cementation is consideredelsewhere (Primmer et al., this volume; Gluyas andColeman, 1992; Gluyas and Witton, this volume) and inother publications (Robinson and Gluyas, 1992; Gluyaset al., 1993a, b), but is outside the scope of this chapter.

DERIVATION OF A POROSITY–DEPTH RELATIONSHIP FOR COMPACTION

We have used two complementary approaches toderive a porosity–depth relationship for compactedsands. The first approach uses results from laboratorycompaction experiments on quartzose sands, andinvolves conversion of the experimentally appliedstress to burial effective stress and depth. The secondapproach uses field data from uncemented sandsaround the world to extend the porosity–depth trend.The close coincidence between the relationshipsderived from the two approaches gives considerableconfidence in their use for porosity prediction.

Approach 1—Using Experimental Data

Most laboratory compaction experiments have acivil engineering or soil science application and areperformed at much lower applied stresses than thoseinvolved in burial to depths in excess of a few hun-dred meters. There are, however, a few examples oftriaxial compression experiments at higher stresses.Vesic and Clough (1968) published the results of testsat loads of ≤30 MPa (4350 psi) on medium-grained,uniform, slightly micaceous quartz sand. Thesestresses can be equated to hydrostatic burial todepths of ~1400 m (4500 ft). Vesic and Clough (1968)also offered a mathematical proof that under mostdeep-burial conditions (<100 MPa), the sand behavedas a linearly deformable solid, with a modulus ofdeformation proportional to the mean normal stress.In other words, the porosity–stress relationship is lin-ear. In addition, Atkinson and Bransby (1978) statethat at high stresses and under what they term nor-mal consolidation conditions (no overpressure),sands will consolidate (compact) so that the relation-ship between incremental applied stress and volumechange/porosity reduction is linear.

The conversion of laboratory pressures and resul-tant porosity values to effective stress or burial depthrelationships raises two important issues. First, doeslaboratory compaction over necessarily short timeperiods (hours or days at most) involve the sameprocesses as burial compaction over much longer geo-logical time periods? Second, how can experimentalstresses can be converted to burial stresses?

Burial Compaction Processes—Laboratory Replication

Sands consisting of quartz grains compact duringburial by a combination of two processes. Theseprocesses are the mechanical response to stress (grainslippage, rotation, and fracturing) and the chemicalprocess of pressure dissolution at grain contacts.Under near-surface conditions, all compaction is bymechanical processes. At a depth of ~1000 m (3000 ft),pressure and temperature become sufficiently ele-vated to permit pressure dissolution (Füchtbauer,1967). From then on, compaction may include a combi-nation of mechanical and chemical processes.

Sands that contain ductile grains in addition toquartz (or other rigid grains) commonly lose porositymore quickly than quartzose sands because the com-pactional process is different (Kurkjy, 1988). Examplesof such grains are mudstone clasts, glauconite grains,phyllitic and schistose metamorphic grains, and micas.Not only do these grains deform more easily andrapidly than rigid grains, but their deformation willpermit a greater degree of slippage and rotation ofmore-rigid grains such as quartz. As a result, a sandwith such grains will lose porosity at a faster rate, par-ticularly during early burial, than a quartzose sand-stone under the same conditions. In addition, allcompaction may be effected by mechanical processes,and pressure dissolution may be relatively unimpor-tant. In this chapter, we present our findings for thecompaction of quartzose (and similar) sands.

Prediction of Porosity in Compacted Sands 21

Page 40: Reservoir Quality Prediction in Sand and Carbonates

22 Gluyas and Cade

Tab

le 1

. Por

osit

y, D

epth

, an

d O

verp

ress

ure

Dat

a fo

r U

nce

men

ted

San

dst

ones

.*

Wel

l/Fi

eld

Dep

thP

oros

ity

Ove

rpre

ssu

re(I

nte

rval

)L

ocat

ion

(km

)(%

)(M

Pa)

Age

Form

atio

nR

efer

ence

Com

men

ts

Frig

gN

OC

S1.

9029

Hyd

rost

atic

Eoc

ene

Frig

gA

bbot

s, 1

991

Fort

ies

UK

CS

2.10

27H

ydro

stat

icPa

leoc

ene

Fort

ies

Abb

ots,

199

1Pi

per

UK

CS

2.40

24H

ydro

stat

icJu

rass

icPi

per

Abb

ots,

199

1D

raug

enN

OC

S1.

6030

Hyd

rost

atic

Jura

ssic

Røg

nB

jørl

ykke

et a

l., 1

986

Abr

oath

UK

CS

2.50

24H

ydro

stat

icPa

leoc

ene

Fort

ies

Abb

ots,

199

1

Bal

mor

alU

KC

S2.

1325

Hyd

rost

atic

Pale

ocen

eA

ndre

wA

bbot

s, 1

991

Mau

reen

UK

CS

2.65

21.5

Hyd

rost

atic

Pale

ocen

eM

aure

enA

bbot

s, 1

991

Mon

tros

eU

KC

S2.

4824

Hyd

rost

atic

Pale

ocen

eFo

rtie

sA

bbot

s, 1

991

Stat

fjord

NO

CS

2.48

27H

ydro

stat

icJu

rass

icB

rent

Gp

Spen

cer

et a

l., 1

987

Stat

fjord

NO

CS

2.55

23H

ydro

stat

icJu

rass

icD

unlin

Gp

Spen

cer

et a

l., 1

987

Stat

fjord

NO

CS

2.70

22H

ydro

stat

icT

rias

sic

Stat

fjord

Spen

cer

et a

l., 1

987

Od

inN

OC

S2.

0029

Hyd

rost

atic

Eoc

ene

Frig

gSp

ence

r et

al.,

198

7N

ord

Øst

Fri

ggN

OC

S1.

9528

Hyd

rost

atic

Eoc

ene

Frig

gSp

ence

r et

al.,

198

st F

rigg

NO

CS

1.90

29H

ydro

stat

icE

ocen

eFr

igg

Spen

cer

et a

l., 1

987

Hei

md

alN

OC

S2.

1025

Hyd

rost

atic

Pale

ocen

eH

eim

dal

Spen

cer

et a

l., 1

987

Slei

pner

Øst

NO

CS

2.33

26H

ydro

stat

icPa

leoc

ene

Hei

md

alSp

ence

r et

al.,

198

7C

odN

OC

S2.

9520

Hyd

rost

atic

Pale

ocen

eFo

rtie

sSp

ence

r et

al.,

198

7G

ullf

aks

NO

CS

1.84

27H

ydro

stat

icJu

rass

icB

rent

Gp

Spen

cer

et a

l., 1

987

Gul

lfak

sN

OC

S1.

9030

Hyd

rost

atic

Jura

ssic

Dun

lin G

pSp

ence

r et

al.,

198

7G

ullf

aks

NO

CS

1.96

25.5

Hyd

rost

atic

Tri

assi

cSt

atfjo

rdSp

ence

r et

al.,

198

7

Tro

llN

OC

S1.

4030

Hyd

rost

atic

Jura

ssic

Sogn

efjø

rdSp

ence

r et

al.,

198

7G

yda

(fie

ld c

rest

)N

OC

S3.

8724

.514

.0Ju

rass

icG

yda

Glu

yas,

199

7Iv

anho

eU

KC

S2.

3924

.6H

ydro

stat

icJu

rass

icPi

per

Park

er, 1

991

Blo

ck 2

2/11

UK

CS

2.40

25.3

Hyd

rost

atic

Pale

ocen

eFo

rtie

sun

publ

ishe

dW

ell 2

2/6a

-2U

KC

S2.

2727

.5H

ydro

stat

icPa

leoc

ene

Fort

ies

unpu

blis

hed

Wel

l 22/

17-1

UK

CS

2.54

22H

ydro

stat

icPa

leoc

ene

Fort

ies

unpu

blis

hed

Wel

l 23/

26a-

2U

KC

S2.

6321

.6H

ydro

stat

icPa

leoc

ene

Fort

ies

unpu

blis

hed

Wel

l 23/

21-1

UK

CS

2.57

23.2

3.0

Pale

ocen

eA

ndre

wun

publ

ishe

dW

ell 2

3/27

-4U

KC

S2.

7522

.53.

7Pa

leoc

ene

Fort

ies

unpu

blis

hed

Page 41: Reservoir Quality Prediction in Sand and Carbonates

Prediction of Porosity in Compacted Sands 23

Wel

l 30/

6-2

UK

CS

2.87

2616

.0Pa

leoc

ene

And

rew

unpu

blis

hed

Gry

phon

UK

CS

1.74

33H

ydro

stat

icE

ocen

eB

ald

erN

ewm

an e

t al.,

199

3M

elib

urSu

mat

ra0.

7537

.5H

ydro

stat

icM

ioce

neSi

hapa

sG

luya

s an

d O

xtob

y, 1

995;

0.

75 k

m =

max

. bur

ial

Gua

poT

rini

dad

1.22

33H

ydro

stat

icPl

ioce

neC

ruse

Bear

d, 1

985;

max

bur

ial ≥

quot

ed d

epth

Fyza

bad

Tri

nid

ad1.

3733

Hyd

rost

atic

Plio

cene

Cru

se &

For

est

Low

, 198

5; m

ax b

uria

l ≥qu

oted

dep

th

Nor

th S

old

ado

Tri

nid

ad2.

0630

11.0

Plio

cene

Man

zani

llaR

adow

sky

and

Iqub

al, 1

985,

as

abov

eC

año

Lim

onC

olom

bia

2.32

24.5

Hyd

rost

atic

Eoc

ene

Mir

ador

McC

ullo

ugh,

199

0M

aui

New

Zea

land

2.74

25H

ydro

stat

icE

ocen

eK

apun

iA

bbot

, 199

0W

ar-W

ink

U.S

.A. (

Tex

as)

4.60

3840

.0Pe

rmia

n–

Luo

et a

l., 1

994;

san

d co

mpo

sitio

n no

t sta

ted

Eas

t Bre

aks

165

U.S

.A. (

GO

M)

3.84

3020

.3Pl

io-P

leis

toce

ne–

unpu

blis

hed

Eas

t Bre

aks

165

U.S

.A. (

GO

M)

3.32

3016

.1Pl

io-P

leis

toce

ne–

unpu

blis

hed

Mis

siss

ippi

U

.S.A

. (G

OM

)3.

0031

15.8

Plio

-Ple

isto

cene

–un

publ

ishe

dC

anyo

n 20

Mis

siss

ippi

U.S

.A. (

GO

M)

1.84

315.

8Pl

io-P

leis

toce

ne–

unpu

blis

hed

Can

yon

20V

iosc

a K

noll

989

U.S

.A. (

GO

M)

2.07

328.

4Pl

io-P

leis

toce

ne–

unpu

blis

hed

Mis

siss

ippi

U

.S.A

. (G

OM

)1.

7533

5.9

Plio

-Ple

isto

cene

–un

publ

ishe

dC

anyo

n 10

9

*NO

CS

= N

orw

egia

n C

onti

nent

al S

helf

; UK

CS

= U

nite

d K

ingd

om C

onti

nent

al S

helf

; GO

M =

Gul

f of M

exic

o.

Page 42: Reservoir Quality Prediction in Sand and Carbonates

24 Gluyas and Cade

The porosity losses sustained during laboratory com-paction experiments occur through grain rearrange-ment, slippage, and fracture (Füchtbauer, 1967) ratherthan pressure dissolution and reprecipitation. As aresult, the experiments may be expected to successfullyreflect compaction in quartzose sands undergoing shal-low burial and in lithic sands (with ductile grains) overmost burial stress conditions. For quartz sands atgreater depths, where chemical compaction processes(e.g., pressure dissolution) become important, theexperiments cannot replicate burial compaction.Despite this, several observations allow the results ofquartz sand compaction experiments to be used. Theexperiments of Vesic and Clough (1968) are restricted torelatively low stresses (burial to ~1400 m) and, over thisdepth range, compaction is predominantly by mechani-cal processes. In addition, porosity–depth data forburied sands show close agreement with the experi-mental results where the two relationships overlap; thelinear consolidation trend described by Atkinson andBransby (1978) matches observed data that show a sim-ilar linear trend at depths in excess of 1000 m (Figure 1).

Laboratory Stress—Conversion to Burial DepthIt is difficult, if not impossible, to exactly replicate the

stress conditions of hydrostatic burial in a laboratory

experiment. To achieve this requires a controlled, zero-stress confinement in all orientations with the exceptionof the vertical, and the application of a unidirectionalvertical stress upon the sample. In addition, pore fluidpressure must be closely controlled to reflect hydro-static fluid pressure increase during burial.

A triaxial test apparatus was used for the laboratoryexperiments of Vesic and Clough (1968). This gives theclosest approximation to true burial. Confinement ofthe sample is controlled by the steel sample chamberand a membrane across which a pressure can beapplied to an oil cushion. Moreover, pore fluid can bedrained from the sample in such a way that pore fluidpressure can be controlled during the experiment. Thesimulation of burial loading is then applied, by apiston, to the top of the sample.

In these experiments, a constant cell pressure wasapplied to the oil cushion surrounding the sample, andpore fluid pressure within the sample was controlled sothat no excess or overpressure (above hydrostatic) wasallowed to develop. Although these triaxial tests do notmimic burial compaction exactly, they do, we believe,give a close approximation. In particular, they do notallow the buildup of overpressure in the pore fluid(which would occur if the pore fluid in the sample wereto be completely confined), nor do they allow pore fluidto drain away freely during compaction, which wouldmean all vertical loading being applied at the grain con-tacts. To equate experimental stresses to burial stresses,the following conversion is used: 1 psi (experiment) =1 ft of burial, or 0.02262 MPa = 1 m of burial.

For the purpose of making the conversion, it isassumed that fluid pressure in the experiments wasmaintained at the equivalent of hydrostatic for theapplied stress. In a normally pressured (hydrostatic)sedimentary sequence, lithostatic load typicallyincreases by 22.622 MPa/km. Using this conversion,the curve of Atkinson and Bransby (1978), usingexperimental data of Vesic and Clough (1968), may beredrawn as a porosity–depth curve for normally pres-sured clean, rigid-grain (quartz, feldspar) sand. Atkin-son and Bransby (1978) do not quote an equation fortheir curve, so an empirical fit has been made to theirexperimentally derived (initially) loose-packed-sandporosity–stress curve.

The equation for this empirical fit is

(1)

where porosity (φ) is in percentages and depth (z) is inmeters.

This equation gives a close fit to the experimentaldata and is in close agreement with field data (Figure 1).In particular, it closely replicates the nonlinear defor-mation behavior of loose sands at low stress. Vesic andClough (1968) performed a number of experimentswith differently packed sand. Starting porosity rangedfrom the 50% used here to 40% for a well-packed sand.However, the initial differences in porosity reduced to~1% at stresses equivalent to 1 km of burial.

φ =+ ×

50

102 4 5 10

3

4exp–

.

–z

z

Figure 1. Squares = porosity to depth relationshipfor hydrostatically pressured (Table 1), uncemented,rigid-grain sandstones. Solid line = porosity todepth relationship for experimental (laboratory)compaction of natural sand under simulated hydro-static conditions. Dashed lines = ±2.5% porosityvariance from experimental curve of Atkinson andBransby (1978).

Page 43: Reservoir Quality Prediction in Sand and Carbonates

Approach 2—Using Field Datafor Uncemented Sands

Porosity data for a group of uncemented buriedsandstones from a range of sedimentary basins havebeen collated (Table 1) and used to produce a predic-tive porosity–depth relationship. Figure 1 shows theexperimentally derived curve (from equation 1) andporosity–depth data from this group of hydrostaticallypressured clean sands. The sands in the group are cur-rently at their maximum burial depth, have hydrostaticpore fluid pressure, and contain <5% ductile grains ordispersed argillaceous material. The close coincidencebetween the curve and the data in the depth range1000–3000 m suggests that the conversion betweenexperimental and burial stresses described in the pre-ceding text is valid. With backup from the real buriedsand data, the new compaction curve for uncementedsandstones provides a reliable means of predictingporosity for uncemented sands. Ninety-five percent ofthe buried sandstone data plotted in Figure 1 fallswithin 2.5 porosity units of the experimentally derived,and extrapolated, porosity–depth curve.

Prediction of Porosity in Clean, Overpressured Sands

In buried sands that have pore fluid pressures sig-nificantly higher than hydrostatic, anomalously highporosities can be preserved. The fluid overpressure(difference between actual pore pressure and hydrosta-tic pressure at the same depth) supports part of theburial loading, and thus reduces the effects of com-paction. Overpressured sands commonly have higherporosity than hydrostatic sands have at the samedepth. Compactional porosity reduction is the result ofeffective stress, which is the difference between litho-static stress (the stress due to the weight of overlyingsediments) and pore fluid pressure. An overpressuredsand will have an effective stress that is equivalent to ahydrostatically pressured sand at a shallower depth(this shallower depth may be termed the “effective bur-ial depth” of the overpressured sand); this depth differ-ence is proportional to the magnitude of theoverpressure. The concept of effective stress and effec-tive burial depth can be used to correct our porosityprediction for an overpressured situation. The effectiveburial depth for an overpressured sand, or the depthunder hydrostatic conditions at which the sand wouldhave the same effective stress, is given by the equation

(2)

where z = burial depth (in meters); z’ = effective burialdepth (in meters); ρr = density of overlying rock col-umn (in Kgm–3) = 2650 (suggested value); ρw = densityof water (in Kgm–3) = 1050 (suggested value); g = grav-ity (in ms–2) = 9.8; φΣ = porosity (as a fraction of 1) = 0.2(suggested value); and u = overpressure (in MPa).

The values given here for density are typical aver-age values for sediments and formation brines. The

suggested value for porosity is a typical figure for theaverage porosity (sands and muds) for a 3-km-thickcolumn of 80% mud and 20% sand. With the sug-gested values, equation 2 works well for burial depthsin the range of 2 to 4 km, and sand to mudstone ratiosof 15:85 to 25:75. For shallow burial depths and/orunusual sand-to-shale ratios, the average porosity ofthe overburden can be calculated by integrating thearea under simple empirically derived porosity–depthfunctions (Baldwin and Butler, 1985).

Equation 2 can be simplified to the following con-versions in order to calculate the effect of 1 MPa(~140 psi) of overpressure in terms of effective depthdifferential. Using the above figures, 1 MPa of over-pressure equates to ~80 m less burial. Thus,

z’ = z – 80u (3)

can be used to derive effective burial depths for substi-tution into the porosity–depth equation.

z zugr w

' –– –

= ( ) ( )

ρ ρ 1 φΣ

Prediction of Porosity in Compacted Sands 25

Figure 2. Open squares = porosity to depth relation-ship for overpressured, uncemented, rigid-grain sand-stones. Solid squares = same sandstones as opensquares but with depth recalculated as effective burialdepth, where effective burial depth (in kilometers) =depth (in kilometers) – 0.08 × overpressure (in MPa).Solid line = porosity to depth relationship for experi-mental (laboratory) compaction of natural sand undersimulated hydrostatic conditions. Dashed lines =±2.5% porosity variance from experimental curve. Foruncorrected data, 33% of predictions fall within the±2.5% range on the mean. This improves to 50% aftercorrection for overpressure. With a more generous dis-tribution about the mean (±5%), prediction accuracyfor overpressure-uncorrected data improves to 50%,while prediction accuracy for overpressure-correcteddata improves to 92%.

Page 44: Reservoir Quality Prediction in Sand and Carbonates

26 Gluyas and Cade

Figure 2 contains a plot of porosity against depth fora variety of overpressured sandstones. A plot of poros-ity against depth for these overpressured sandstones,but with their depths adjusted to effective burial depthusing equation 2, is also shown in Figure 2. There isclose agreement between the experimentally derivedporosity–depth trend and the measured porosity/effec-tive burial depth data for the overpressured sandstones.

DISCUSSION

We believe that the porosity–depth relationship pre-sented in this chapter gives the explorer a valuable toolfor the prediction of sandstone porosity ahead ofdrilling. The previously published global porosity–depth curves carry too much uncertainty for uses otherthan the prediction of average behavior of a sand underburial. The compaction equations presented here give

an understanding of how porosity is lost as a functionof effective stress. This allows the effect of overpres-sure as well as burial depth to be accounted for. Over-pressured sands are generally more porous than theirhydrostatically pressured counterparts. The porosityof overpressured sands is not accurately predicted bysimple empirical porosity–depth relationships. TheCretaceous Tuscaloosa sandstone of the U.S. GulfCoast area (Thomson, 1979) is a prime example ofporous sandstone at depth that would have been pre-dicted as having insufficient porosity for commercialflow rates, if a simple porosity–depth function fromother Cretaceous sandstones in the area (Figure 3) hadbeen used. However, had equations 1 and 3 been usedalong with an overpressure estimate, the maximumpredicted porosity at 6.4 km would have been esti-mated to be 30%, rather than 2%, from the empiricalrelationship (compared with actual porosity of 23.5%minus cement porosity of 30%).

CONCLUSIONS

Porosity–depth functions are the most commonmethod used for the prediction of sandstone porosityahead of exploration drilling. Where they includesandstones with varying degrees of compaction,cementation, or overpressure, they will often carry alarge range of uncertainty.

Experimental data on the relationship between sand-stone porosity and confining stress provide the explo-ration geoscientist with an alternative method forpredicting porosity at depth in exploration prospects.The equations presented in this chapter are derivedfrom experimental data and have been tested against adiverse worldwide set of buried-sand data. They allowprediction of porosity to ±2.5 porosity units (at 95% con-fidence levels) for clean, normally pressured, uncemented sands. Moreover, through the link betweenporosity and effective stress, the equations deliver amethodology that allows prediction of anomalousporosity preservation due to the effects of overpressure.

ACKNOWLEDGMENTS

We thank BP Exploration for permission to publishthis paper. We also thank Mike Bowman, David Epps,Shona Grant, Nick Milton, Steve Franks, and JohnAggett for their thorough and constructive reviews.

REFERENCES CITED

Abbot, W.O., 1990, Maui field, in E.A. Beaumont and N.H.Foster, eds., Structural traps I: AAPG Treatise ofPetroleum Geology, Atlas of Oil and Gas Fields, p. 1–25.

Abbots, I.L., 1991, United Kingdom oil and gas fields, 25years commemorative volume: Geological Society ofLondon Memoir 14, 573 p.

Athy, L.F., 1930, Density, porosity and compaction ofsedimentary rocks: AAPG Bulletin, v. 14, p. 1–24.

Atkinson, J.H., and P.L. Bransby, 1978, The mechanics ofsoils: an introduction to critical state soil mechanics:London, McGraw Hill, 375 p.

Figure 3. Porosity–depth plot for CretaceousTuscaloosa sandstones of Louisiana (open squares;Thomson, 1979), including those from AlmaPlantation field (point c, solid squares). Simpleregression of porosity on depth = a. Compactioncurve from this chapter (point b). The Tuscalooscasandstones at >6 km typically require mud weightsof 16–17 lb/gal (1.59–1.69 kg/L) (Gill, 1980); equal to~8500–9000 psi (~60 MPa) overpressure. From equa-tion 2, this is 4.8 km less than the actual depth of 6.4km (d). Using equation 1, we would predict a poros-ity of ~30% for the Tuscaloosa sandstones fromAlma Plantation field compared with an actual aver-age porosity for the sands of 23.5% (star). The factthat the Alma Plantation sandstones are partiallycemented by chlorite has been ignored in this calcu-lation. Thomson (1979) quotes 30% cement porosity(point e) for these sandstones; that is exactly as pre-dicted from our compaction equation.

Page 45: Reservoir Quality Prediction in Sand and Carbonates

Baldwin, B., and C.O. Butler, 1985, Compactioncurves: AAPG Bulletin, v. 69, p. 622–626.

Beard, J.T., 1985, The geology of the Guapo field, in B.Carr-Brown and J.T. Christian, eds., Transactions ofthe 4th Latin American Geological Congress,Trinidad and Tobago 1979: Arima, Trinidad &Tobago Ltd, July 7–15, 1979, Port of Spain, p. 684–689.

Bjørlykke, K., P. Aagaard, D. Dypvik, D.S. Hastings, andA.S. Harper, 1986, Diagenesis and reservoir proper-ties of the Jurassic sandstones from the Haltenbankenarea, offshore mid-Norway, in Proceedings of theNorwegian Petroleum Society, Symposium on Habi-tat of Hydrocarbons—Norwegian Oil and Gas Finds:Stavanger, Norwegian Petroleum Society, p. 275–286.

Füchtbauer, H., 1967, Influence of different types ofdiagenesis on sandstone porosity, in W. Ruhl, ed.,Proceedings of the 7th World Petroleum Congress,Mexico: Mexico City, vol. 2, p. 353–367.

Gill, J.A., 1980, Multiparameter log tracks, TuscaloosaWoodbine pressures (abs.): Oil and Gas Journal,November, 3, p. 20–22.

Gluyas, J.G., this volume, Poroperm prediction forreserves growth exploration: Ula Trend, NorwegianNorth Sea, in J. Kupecz, J.G. Gluyas, and S. Bloch,eds., Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 201–210.

Gluyas, J.G., and M.L. Coleman, 1992, Material fluxand porosity changes during diagenesis: Nature,v. 356, p. 52–53.

Gluyas, J.G., and N.H. Oxtoby, 1995, Diagenesis ashort (2 million year) story—Miocene sandstones ofcentral Sumatra, Indonesia: Journal of SedimentaryResearch, v. A65, p. 513–521.

Gluyas, J.G., A.G. Robinson, D. Emery, S.M. Grant,and N.H. Oxtoby, 1993a, The link betweenpetroleum emplacement and sandstone cementa-tion, in J.R. Parker, ed., Petroleum geology of NWEurope: London Geological Society Publication,Proceedings of 4th Conference, p. 1395-1402.

Gluyas, J.G., A.G. Robinson, and S.M. Grant, 1993b,Geochemical evidence for a temporal control onsandstone cementation: AAPG Studies in Geology36, p. 23–33.

Gluyas, J.G., and T. Witton, this volume, Porosity andpermeability prediction for wildcat explorationdrilling, Miocene Southern Red Sea, in J. Kupecz, J.G.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir 69,p. 163–176.

Kurkjy, K.A., 1988, Experimental compaction studies oflithic sands: M.S. thesis, Rosensteil School of Marineand Atmospheric Sciences, University of Miami, 101 p.

Low, B.M., 1985, The geology of the Fyzabad main field,in B. Carr-Brown and J.T. Christian, eds., Transac-tions of the 4th Latin American Geological Congress,Trinidad and Tobago:Arima, Trinidad & TobagoLtd, July 7–15, 1979, Port of Spain, p. 714–719.

Luo, M., M.R. Baker, and D.V. LeMone, 1994, Distribu-tion and generation of the overpressure system,

eastern Delaware Basin, western Texas and southernNew Mexico: AAPG Bulletin, v. 78, p. 1386–1405.

McCullough, C.N., 1990, Caño Limon field, LlanosBasin, Colombia, in E.A. Beaumont and N.H. Foster,eds., Structural traps II: AAPG Treatise of PetroleumGeology, Atlas of Oil and Gas Fields, p. 65–93.

Newman, M.St.J., M.L. Reeder, A.H.W. Woodruff, andI.R. Hatton, 1993, The geology of the Gryphon oilfield, in J.R. Parker, ed., Petroleum geology of NWEurope: London Geological Society Publication,Proceedings of 4th Conference, p. 123–133.

Parker, R.H., 1991, The Ivanhoe and Rob Roy fields,Block 15/21a-b, UK North Sea, in I.L. Abbots, ed.,United Kingdom oil and gas fields, 25 years com-memorative volume: Geological Society of LondonMemoir 14, p. 331–338.

Pettijohn, F.J., 1975, Sedimentary rocks (3d ed.): NewYork, Springer-Verlag, 628 p.

Primmer, T.P., C.A. Cade, I.J. Evans, J.G. Gluyas, M.S.Hopkins, N.H. Oxtoby, P.C. Smalley, E.A. Warren,and R.H. Worden, this volume, Global patterns insandstone diagenesis: their application to reservoirquality prediction for petroleum exploration, in J.Kupecz, J.G. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 61–78.

Radowsky, B., and J. Iqubal, 1985, Geology of the NorthSoldado field, in B. Carr-Brown and J.T. Christian,eds., Transactions of the 4th Latin American Geologi-cal Congress, Trinidad and Tobago 1979; Arima,Trinidad & Tobago Ltd, July 7–15, 1979, Port of Spain,p. 759-769.

Ramm, M., A.W. Forsberg, and J. Jahren, this volume,Porosity depth trends in deeply buried Upper Juras-sic reservoirs in the Norwegian Central Graben: anexample of porosity preservation beneath the normaleconomic basement by grain-coating micro-quartz, inJ. Kupecz, J.G. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 177–200.

Robinson, A.G., and J.G. Gluyas, 1992, Model calcula-tions of sandstone porosity loss due to compactionand quartz cementation: Marine and PetroleumGeology, v. 9, p. 319–323.

Scherer, M., 1987, Parameters influencing porosity insandstones: a model for sandstone porosity predic-tion: AAPG Bulletin, v. 75, p. 485–491.

Selley, R.C., 1978, Porosity gradients in North Sea oil-bearing sandstones: Journal of the Geological Soci-ety of London, v. 135, p. 119–132.

Spencer, A.M., et al., 1987, Geology of the Norwegian oiland gas fields: Stavanger, Graham and Trotman, 443 p.

Thomson, A., 1979, Preservation of porosity in the deepWoodbine-Tuscaloosa trend, Louisiana: Gulf CoastAssociation of Geological Society Transactions, v. 30,p. 396–403.

Vesic, A.S., and G.W. Clough, 1968, Behaviour of gran-ular material under high stresses: Journal of SoilMechanics Foundation Division, v. 94, p. 661–688.

Weller, J. M., 1959, Compaction of sediments: AAPGBulletin, v. 43, p. 273–310.

Prediction of Porosity in Compacted Sands 27

Page 46: Reservoir Quality Prediction in Sand and Carbonates

29

Chapter 3

Porosity Variation in Carbonates as aFunction of Depth: Mississippian Madison

Group, Williston BasinAlton Brown

ARCO Exploration and Production Technology CompanyPlano, Texas, U.S.A.

ABSTRACT

Log-determined porosities of argillaceous limestone, limestone, dolomiticlimestone, and dolomite of the Mississippian Madison Group in theWilliston Basin were analyzed to determine the influence of carbonate min-eralogy, shale content, and fabric on porosity loss with depth of burial.Carbonate mineralogy and shale content strongly influence the rate of poros-ity loss. Argillaceous carbonates lose porosity at the greatest rate with burial,followed by clean limestone, dolomitic limestone, and dolomite. Averageporosity of grain-supported limestone is not systematically higher than aver-age porosity of mud-supported limestone in the same depth range, but thereis a significant difference in the respective porosity range. Moderately todeeply buried (1.5–3 km) limestones with a grain-supported texture have asmall percentage of high-porosity samples, whereas porosity distributions inmatrix-supported limestones at equal burial depth cluster around the meanporosity and lack a tail of high-porosity samples. This effectively limits eco-nomic porosity in moderately to deeply buried Madison limestones to grain-supported rocks (packstones and grainstones).

Results of this study reveal characteristics of basin-scale porosity lossmechanisms. Secondary porosity formed during burial is not evident in theporosity–depth profiles. Porosity loss is strongly influenced by mineralogy;clay content greatly accelerates the rate of porosity loss in limestones. In theserocks, dolomite porosity higher than limestone porosity at a given maximumburial depth is due primarily to selective preservation of dolomite porosity.Porosity decreases with increasing temperature in rocks with otherwise simi-lar burial (effective stress) history. The observed porosity–depth relationshipsroughly follow an exponential trend; this may indicate that there is some sortof feedback between porosity and the porosity reduction mechanism.

Brown, A., 1997, Porosity variation in carbonates as afunction of depth: Mississippian Madison Group,Williston Basin, in J.A. Kupecz, J. Gluyas, and S.Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69, p. 29–46.

Page 47: Reservoir Quality Prediction in Sand and Carbonates

30 Brown

INTRODUCTION

Porosity in sediments is strongly influenced by depo-sitional and early diagenetic environments. Upon burial,porosity is usually lost as pressure, temperature, andtime of exposure to diagenetic environments increase.The path of the porosity loss is of major economicimportance because the path determines the distributionof porosity in buried rocks, and directly influences thelikelihood of economic hydrocarbon accumulations.

Porosity in chalks and deep-water limestones haslong been recognized to be predominantly influencedby depth of burial (Scholle, 1978). However, early dia-genesis has long been believed to have a much greatereffect on porosity in shallow-water carbonate rocksthan burial diagenesis (Choquette and Pray, 1970).Subsequent work has demonstrated that burial is amajor control on average porosity distribution inshallow-water limestones in some basins (Schmokerand Halley, 1982; Schmoker, 1984).

The observed correlations between depth of burialand porosity raise many questions. What are the possi-ble effects of sampling strategy on porosity–depthtrends? What is the effect of lithology types on porosityloss with burial? What are the effects of depositional tex-ture on porosity loss? How can one evaluate the tremen-dous scatter characteristic of shelf limestone porosity vs.depth data in order to make predictions about the likeli-hood of encountering economic porosity?

In an attempt to answer some of these questions,porosity data from wells penetrating the MadisonGroup (Mississippian) carbonates of the Williston Basinwere analyzed for correlations to lithology, deposi-tional fabric, depth, and temperature gradient. Results

presented here document the strong effect of carbonatemineralogy and relatively weak effect of depositionaltexture on average porosity trends with depth of burial.The major effect of depositional texture is preservationof a broader distribution of porosity in grain-supportedlimestone textures than in mud-supported limestonetextures in a given depth range.

PREVIOUS WORK

Previous studies of porosity–depth relationships inshallow-water carbonates evaluated two types of data:measurements of reservoir porosity of petroleumaccumulations (Schmoker et al., 1985), and measure-ments of porosity of the carbonate on a basinal scale(Schmoker and Halley, 1982).

Reservoir porosity studies such as those done bySchmoker et al. (1985) provide a basis for evaluatingthe properties of discovered and undiscovered eco-nomic reservoirs. These types of data provide esti-mates of the expected porosity and other reservoircharacteristics of fields once they are discovered.Reservoir properties may or may not reflect the prop-erties of the basin-scale carbonate, because for a reser-voir to be economic, it must have some minimumreservoir quality. The substantial fraction of carbon-ates that have low porosity are not represented in thefield databases. This means that the porosity distribu-tion of reservoirs may not indicate the likelihood ofencountering economic porosity during wildcat explo-ration, but it may indicate the porosity likely to befound in economic discoveries.

Basin-scale studies of shallow-water carbonateporosity–depth relationships initiated with the

Data generated in this study can be used to predict porosity distribution at agiven depth in the Mississippian strata of the Williston Basin if no other infor-mation is available. Average limestone porosity at moderate to deep burial issignificantly less than the porosity required for economic development ofunfractured petroleum accumulations, so average porosity cannot be used asan estimate of economic porosity in a prospect. However, the distribution ofporosity in a depth range can be used to estimate the risk associated withencountering sufficient thickness of economic porosity. The presence or absenceof potentially economic porosity is best evaluated as a risk statement. For thisreason, the porosity cumulative frequency distribution in a given depth rangeis a particularly useful tool because it can be interpreted in terms of expectedthickness of porosity higher than a given threshold value. If information aboutvertical spatial correlation of porosity is available, the distribution can be inter-preted in terms of risk of finding a minimum net thickness of carbonate exceed-ing a threshold porosity level. These methods can be used in other wildcatexploration settings where proper calibration data have been collected. Theresults of this study can be used as a guide to understanding porosity distribu-tion with depth in other Paleozoic carbonates, and perhaps be directly appliedto other late Paleozoic carbonates in cratonic settings.

Page 48: Reservoir Quality Prediction in Sand and Carbonates

now-classic study by Schmoker and Halley (1982).Because all carbonates were sampled in the approachused by those authors, conclusions based on their dataapply to basin-scale trends of porosity development.They clearly demonstrated the systematic decrease ofporosity with burial depth, evident not only in interme-diate to deep samples (Schmoker and Halley, 1982), butalso in shallow samples (Halley and Schmoker, 1983).These data also document the different porosity-losspathways of dolomite and limestone, and the approxi-mately exponential shape to the trend of porosity loss(Schmoker and Halley, 1982). A subsequent study bySchmoker (1984) evaluated a number of carbonateporosity–depth trends and found that, in general, theyfollowed a log-linear relationship to time-temperatureindex (TTI), a measure of thermal exposure.

Two problems crop up in the previous studies: (1)using reservoir data to characterize basinal porositytrends and (2) possible bias in selection of basin-scaledata. As long as the differences between the uses ofreservoir data and basinal data are recognized, no con-fusion results. Just as reservoir data cannot be used tocharacterize the basin-scale changes in limestonereservoir quality in an unbiased manner, the basin-scale data analyses include data that have uneconomicporosity, so averages of these data do not reflect theporosity of expected discoveries. This distinction hasnot always been clear in previous studies. Schmoker(1984) used reservoir data in order to characterizebasin-scale carbonate properties in addition to otherbasin-scale data. Although this does not invalidate hisresults, some of the scatter in the trends of porosity toTTI may be explained by the use of reservoir data setsto characterize a basin-scale process.

Some previous basin-scale studies used data collec-tion techniques that can introduce a bias of unknownmagnitude. For example, the sampling approach ofSchmoker and Halley (1982) introduced a bias to theirdata. They measured average porosity of intervals withrelatively constant porosity. This approach does notmeasure porosity on a volumetric basis, because ashort interval of low porosity carries as much weight-ing as a much longer interval of high porosity. As aver-age porosity decreases with depth, the high-porosityintervals generally become shorter and the low-porosity intervals become longer. This means that theshallow intervals may be systematically biased towardlow-porosity, whereas the deeper intervals may bebiased toward high porosity. Although interval lengthvaries by a factor of 7 in their data, the variable intervallength has not introduced enough error to invalidatethe conclusions of Schmoker and Halley (1982). Quan-titative use of this data set for testing models of poros-ity loss may be affected by this bias, however.

This brief review indicates why this study wasundertaken in the manner it was. The main goal of thestudy was to evaluate basin-scale trends of porosityevolution. This requires careful consideration of thesampling strategy in order to collect an unbiased,basin-scale database. The secondary goal is developingmethods of quantitative prediction of wildcat risk forreservoir quality. It is believed that the only successfulstrategy for predicting economic porosity is to considerthe distribution of all porosity within the interval of

interest: uneconomic porosity as well as potentiallyeconomic porosity levels.

STUDY AREA AND METHODS

Setting

The Madison Group of the Williston Basin was cho-sen for study for the following reasons. (1) The samegeneral stratigraphic interval could be sampled at var-ious depths of burial. By sampling rocks of a narrowage range, time effects on porosity loss can be mini-mized. (2) A large number of well logs with modernporosity logging packages are available over a largegeographic area. This removes possible geographicbias. (3) Modern depths of burial in the study area ineastern and central Williston Basin are close to maxi-mum burial experienced by the basin, although therehas been minor Cenozoic erosion around the marginof the basin. Thus, present subsurface temperaturesare probably close to the maximum temperatures towhich the carbonates were exposed. Williston Basinsubsidence is somewhat episodic, but samples fromdifferent burial depths have very similar relative sub-sidence curves (Figure 1). This means that differencesin burial history are not likely to affect porosity evolu-tion. (4) A variety of carbonate mineralogies and tex-tures is present in the Madison Group. Mineralogiescan be identified from log analysis due to the relativelysimple mineralogical composition of the carbonates.This allows for accurate porosity determination for arange of carbonate lithologies. Generalized carbonatefabric data are available from cuttings descriptions.

The Madison Group is a Kinderhookian to Merime-cean, argillaceous carbonate, carbonate, and evaporiteunit that in fills the Williston Basin by progradationfrom east and south North Dakota (Peterson and Mac-Cary, 1987). The group shoals upward from argilla-ceous limestone and shale near the base (LodgepoleLimestone) through interbedded carbonates andanhydrite (Mission Canyon Formation) to salina salts(Charles Formation) at the top of the group. Faciestracts generally prograde to the west and northwest,resulting in distribution of nearly all major lithofaciesover essentially all parts of central and western NorthDakota, although the facies tracts are not exactly con-temporaneous.

Data Collection and Analysis

Data were collected from the Lodgepole and MissionCanyon formations of the Madison Group. Thirty-onewells were selected to sample the Madison Group at arange of depths and geographic areas in western andcentral North Dakota (Figure 2, Table 1). Porosities andmineralogies were determined from wireline logs. Digi-tized wireline well logs were not available, so lithologyand porosity were determined by manual cross-plottingtechniques of data from paper copies. To create a bias-free data set, porosity and lithology were collected atexact 3.3-m (10-ft) depth marks on wireline well logs,starting from the base of the Charles Salt through thebase of the Lodgepole limestone. Only carbonate litholo-gies were analyzed; shale units in the LodgepoleFormation and the evaporite beds were not evaluated.

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 31

Page 49: Reservoir Quality Prediction in Sand and Carbonates

32 Brown

If the log readings were unreliable at the 10-ft depthmark due to borehole conditions or bed edge effects,the interpretations were made at a depth of 0.6 m (2 ft)above the 10-ft mark. Because of the equal and arbitrar-ily spaced sampling interval, the sample set providesan unbiased estimate of the different lithologies andporosity in the rocks.

Necessary borehole corrections were made beforemineralogical and porosity evaluation from wirelinelog readings. Argillaceous carbonates were identifiedby high gamma-ray (GR) response (>30° API unitsafter mud weight and caliper correction) combinedwith elevated neutron log porosity and depressedsonic log response. Porosity of argillaceous limestoneswas interpreted from compensated density logs, usinga grain density of 2.71 g/cm3. Density porosity esti-mates are relatively insensitive to changes in matrixmineralogy in this setting because the matrix density oflimestone is similar to that of the silicate minerals. Inthese rocks, the predominant silicate mineral is illite,which has a density of 2.77 g/cm3 (Ellis et al., 1988),somewhat higher than 2.71 g/cm3 assumed in theporosity model. Also, small quantities of pyrite (graindensity of 5.0 g/cm3) are routinely reported in descrip-tions of cuttings of the argillaceous limestones in thestudied wells. These compositional differences can leadto an actual matrix density slightly higher than theassumed 2.71 g/cm3, resulting in a small systematic biasfor argillaceous samples toward low porosity.This biasis thought to be <3% in the worst case. Reported nega-tive porosity probably represents in-situ porosity <1%combined with a matrix density >2.71 g/cm3.

Limestones, dolomitic limestones, and dolomiteswere distinguished by compensated density log–compensated neutron log crossplots. The plots also

provided porosity estimates. Because the wells hadslightly different porosity logging tools, differentcharts were used on the wells as appropriate. Overintervals with questionable mineralogy (such as halite-or anhydrite-cemented limestone), cuttings descrip-tion, compensated sonic tool response, and resistivitytool responses were used to confirm mineralogy.Halite- and anhydrite-cemented limestones were notincluded in the data set.

In addition to the porosity and lithology information,texture, temperature, and effective stress were estimatedfor all of the depth intervals for which porosity was mea-sured. Depositional fabrics were estimated from com-mercial sample logs provided by the AMSTRAT(American Stratigraphic) Company. Most wells hadAMSTRAT cuttings logs; those that did not hadAMSTRAT cuttings logs available within a few miles ofthe analyzed wells (Figure 2). These descriptions, inmost cases, could be readily correlated to the studywells. Some wells lack textural data due to lack of cut-tings descriptions. Dolomites had poor description ofdepositional texture. Argillaceous limestones wereinvariably described as mudstones or wackestones. Forthese reasons, the effect of texture on dolomite,dolomitic limestone, and argillaceous limestone poros-ity was not investigated. AMSTRAT cuttings descrip-tions were used to group limestones by texture into fourtextural classes (Dunham, 1962): mudstones, wacke-stones, packstones, and grainstones. Boundstones werenot observed in the Madison Group. Because so fewgrainstones were described, packstones and grainstoneswere combined into the single category of grain-supported rocks for some analyses.

Thermal gradient varies significantly over theWilliston Basin and is apparently not directly related

Figure 1. Comparison of burial history between a basin center well (#7 in Table 1) and a basin margin well (#32in Table 1). The burial depth is scaled in percentage of present-day burial. The major difference in relativeburial history is the preservation of the late Paleozoic–early Mesozoic age strata in the basin center and itsabsence in the basin margin. The basin margin well has also been exhumed somewhat more than the basincenter well, but the exhumation in both cases probably does not exceed 300 m. The similarity of burial histories indicates that the porosity changes correlate to relative burial depths, not differences in burial history.Figured burial curves are constructed from undecompacted formation thicknesses.

(Ma)

Page 50: Reservoir Quality Prediction in Sand and Carbonates

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 33

Figu

re 2

. Loc

atio

ns

of s

tud

ied

wel

ls w

ith

AM

ST

RA

T d

ata

(fil

led

cir

cles

), w

ith

nea

rby

corr

elat

ive

AM

ST

RA

T d

ata

(hal

f-fi

lled

cir

cles

),an

d w

ith

no

text

ura

l dat

a (e

mp

ty c

ircl

es).

Con

tou

rs a

re d

epth

in f

eet t

o th

e to

p o

f th

e L

odge

pol

e Fo

rmat

ion

of

the

Mad

ison

Gro

up

fro

mth

e su

rfac

e.

Page 51: Reservoir Quality Prediction in Sand and Carbonates

34 Brown

to the amount of basin subsidence (DeFord et al.,1976). The differences in thermal gradient can be usedto determine the effects of temperature differences onporosity loss with depth. As the carbonate in each wellis buried, temperature increases, so that the presenttemperature is only the most recent part of a rock’sthermal history. For this reason, thermal gradientsrather than present temperatures are used to distin-guish temperature populations. The wells analyzedwere divided into three groups by the thermal gradi-ent present within the well: high gradient (>31°C/km;>1.7°F/100 ft), moderate gradient (25.5°–31°C/km;l.4–1.7°F/100 ft), and low gradient (<25.5°C/km;<1.4°F/100 ft). Because the surface temperatures ofthese wells are almost identical, the ranking by ther-mal gradient also ranks the wells by temperature atany depth. Temperature differences between the wellsin high thermal gradients and low thermal gradientsare about 22°C at 3 km (40°F at 10,000 ft burial).

The pore pressure of analyzed Mississippian carbon-ates was estimated from the potentiometric surface mapof Miller and Strauz (1980). The present-day verticaleffective stress was calculated by subtracting the porepressure from the geostatic load. It is assumed that the

vertical weight of overlying strata is the maximumstress on the rock, so the vertical effective stress is themaximum effective stress.

The porosity, texture, temperature, effective stress,and depth data were analyzed with the SAS (StatisticalAnalysis Services, Inc.) mainframe statistical package.The porosity distribution was evaluated in severalways. First, samples were grouped into arbitrarydepth intervals of 152 or 305 m (500 or 1100 ft),depending upon sample density. Porosities of differ-ent mineralogies and textures were then averagedover the depth interval. The porosity averaged bylithology or texture was then plotted against depth.The advantage of this approach over a depth-regres-sion model is that depth intervals with high samplesize do not influence the porosity estimated for depthintervals with smaller sample size. Variation of poros-ity with depth is also not constrained to a particularfunctional form (such as linear and exponential).

Porosity was also evaluated by plotting the depthaverages of porosity from wells with the thermal gradi-ent ranges discussed above. The significance of tempera-ture and other variables in an overall regression modelwas also evaluated by SAS GLM, a general linear

Table 1. Studied Wells.

No. Well Location

1 FUCE Jayhawk-Nelson # 43x-30 sec. 30 T33N/R56E (MT)*2 Gulf Lee Mae #1-33-1a sec. 33 T33N/R58E (MT)3 ARCO Wunderlich #1 sec. 22 T151N/R80W4 Amoco Sondrol #1 sec. 10 T149N/R99W5 ARCO Klain #1 sec. 26 T149NR80W7 Placid Rosenthal # 36-5 sec. 36 T163N/R80W8 Cities Service Rice #1 sec. 27 T161N/R82W9 Asmera Welch #1 sec. 31 T138N/R78W10 Gas Prod. Enterprise BN #1 sec. 27 T132N/R86W11 Energetics, Inc. Soelberg #23-7 sec. 7 T130N/R91W12 Tenneco #1-1 Reistad sec. 1 T162N/102W13 Amoco Richter #1 sec. 26 T140 N/R88W14 Mitchell - Elberg # 1-35 sec. 35 T152N/R90W15 Gulf Juma #1-1-1D sec. 1 T156N/R92W16 Hunt Barta #1 sec. 5 T140N/R95W17 Hunt - Treffry #1 sec. 30 T155N/R100W18 Shell USA #22-24 sec. 24 T148N/R103W19 Texaco Luttin #1 sec. 27 T151N/R99W20 Shell Quinell #31-14 sec. 14 T146N/R104W21 Getty Vetter #1 sec. 34 T152N/R73W22 AMOCO Karch #1 sec. 6 T138N/R85W23 Shell Lindbad #41-16 sec. 16 T163N/R87W24 Adobe Oil 23-31X Luptak sec. 31 T141N/R99W25 Pennzoil Railroad Bend #2-24 sec. 2 T134N/R83W26 Terra Resources Borth #1-35 sec. 35 T145N/R93W27 Gulf Rough Rider Federal #1-21-30 sec. 21 T145N/R100W28 Conoco Entze #29-1 sec. 29 T144N/R90W29 Amoco Thompson #8-1a sec. 29 T144N/R99W30 Amoco J. Christman #1 sec. 28 T130N/R95W31 Amoco Kenny # 1 sec. 14 T136N/R96W32 Supron Privatsky #1 sec. 26 T138N/R98W

*MT = Montana.

Page 52: Reservoir Quality Prediction in Sand and Carbonates

model similar to ANOVA. This was used to determinethe statistical significance of class (discontinuous) vari-ables such as the well number and texture, as well as thecontinuous variables of depth and effective stress, on theaverage porosity.

The effect of shale content on limestone porosity wasof particular interest, because clay content is proposedto enhance pressure dissolution (Weyl, 1959) and, there-fore, to decrease porosity by cementation in the burialenvironment. The interval chosen for this study is theLodgepole Formation, which has argillaceous carbon-ate mudstones and wackestones interbedded withshale-free carbonate mudstones and wackestones. Therelationship between porosity and shale content is besttested by comparing the density porosity to a shalinessindicator within single boreholes. Based on cuttingsdescription, G-R intensity increases with increasingshale content and is therefore a suitable shaliness indi-cator for these rocks. Four of the study wells (listed inTable 2) were digitized, the G-R intensity corrected forborehole effects, and the corrected density porosityplotted as a function of G-R intensity in each well.Effects of shale content were then evaluated by linearregression for each of the four wells (Table 2).

Average porosity data for Madison Group petroleumaccumulations were also compiled along with the domi-nant mineralogy of the reservoir rock. As noted by Lind-say (1985), reservoirs north and east of the basin centerare predominantly limestone, whereas those to thesouth and west are predominantly dolomitized or par-tially dolomitized reservoirs. Reservoir lithology andporosity data were compiled from Tonnesen (1985),Tyler (1962), and field papers by Kupecz (1984), Lindsayand Kendall (1985), LeFever and LeFever (1991), Beachand Griffin (1992), and DeMis (1992). The data set is lim-ited in two ways. First, average porosity in fracturedlimestone traps (such as Mondak field) is not reported infield papers, so the average reservoir porosity compiledhere is believed to represent matrix porosity averages.Second, some of the field descriptions in the compilationvolumes (Tyler, 1962; Tonnesen, 1985) report a lime-stone porosity, whereas examination of well logs indi-cates a dolomitic limestone reservoir. For this reason, notall reservoirs reported as limestone in this compilationmay actually be limestone.

RESULTSEffect of Carbonate Lithology

Mean porosity decreases as a function of depth for alllithologies investigated (Figure 3). The porosity–depth

trends decrease in a manner quite similar to exponentialporosity loss with depth, as can be seen where the poros-ity is plotted on a logarithmic scale (Figure 4). Figure 5shows mean, standard deviation, and maximum poros-ity for each depth interval and lithology type.

Clay-free limestones (those with <30° API units cor-rected G-R response) show a systematic, gradualdecrease of porosity with depth (Figure 5a). Thesmoothness of the porosity decrease probably reflectsthe large sample size for clean limestones comparedwith other carbonate lithologies sampled. Dolomiticlimestones (Figure 5c) and dolostones (Figure 5d) havea more erratic porosity decrease with increasing depth.This probably reflects the small sample size at eachdepth interval and the more complex diagenetic historyof dolomitization. At least two types of dolomite wereincluded in the samples of this study. Dolomite in somewells is the high G-R marker bed dolomites, whereas inothers, the dolomite is the product of more pervasivedolomitization. Kupecz (1984) also reports that high-porosity dolomite is associated with petrographic evi-dence of anhydrite secondary dissolution. Porosity ofthe dolomitic limestone samples is affected by both sec-ondary anhydrite dissolution and by variable fractionof dolomitization. The rate of porosity loss of dolomitewith increasing depth is only slightly less than that ofdolomitic limestone. The major difference between thetwo trends is that dolomite has an average porosityhigher than dolomitic limestone at the shallowest sample depths.

Argillaceous limestones have the lowest porosity atany depth (Figure 3). Porosity decreases rapidly to 3%at 1.5–2.6 km (5000–8500 ft), and then decreasesslowly with increasing depth (Figure 5b). The samplesfrom depths to 2.6–3.0 km (8500–10,000 ft) have anaverage density porosity that is negative. The nega-tive density porosity of the argillaceous limestonesprobably represents porosity <1% combined with amatrix density greater than the assumed 2.71 g/cm3,as discussed above.

Effect of Clay Content

In the three wells with samples of shallow to interme-diate depth (<2 km; <6000 ft), density porosity systemat-ically decreases with increasing G-R intensity (Figure 6a;Table 2). The observed decrease in density porosity withincreasing shale content cannot be fully explained byany reasonable change in matrix density; therefore, thesystematic porosity decrease is interpreted to be causedprimarily by increasing clay content.

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 35

Table 2. Porosity Correlations to Gamma-Ray Intensity.

Depth Investigated Regression Zero G-RWell # (ft) Slope* Intercept * R2

7 3800–4100 –0.00188 0.142 0.588 4540–4880 –0.0024 0.128 0.493 5300–5700 –0.00178 0.07 0.664 10,180–10400 –0.00012 –0.020 0.026

* Least-squares model between fractional porosity and gamma-ray (G-R) intensity in API units.

Page 53: Reservoir Quality Prediction in Sand and Carbonates

36 Brown

The well with deeply buried (~3.3 km; 11,000 ft)samples shows no significant correlation between den-sity porosity and GR intensity (Table 2; Figure 6b). Thelack of correlation is believed to be caused by the rela-tively small range of porosity in the evaluated depthinterval. This is consistent with the regional patternshown in Figure 3; namely, clean limestones have verylow porosity at this depth, so any further decrease inporosity with increasing clay content would not besignificant. Much of the variation in log-determinedporosity in Figure 6b is probably caused by variationsin matrix density; this is the reason that many of theporosity measurements have negative values.

The intercept of the regression equations in Table 2theoretically represents the porosity of a limestonewith a GR intensity of zero. The intercept porosities forthe first three wells in Table 2 are significantly higherthan porosities shown for clean limestones at compa-rable depths in Figures 3 and 5. The difference iscaused by the fact that almost all limestones have a GRresponse >10° API units. Clean limestones can have≤30° API units of corrected GR. Porosity calculatedfrom the regression equations using the actual valuesof GR intensity is within the range of average porosityshown on Figure 3.

Effect of Texture

The general linear regression model demonstratesthat fabric has a significant effect on limestone poros-ity; this effect is much less significant than the effect ofdepth and that of effective stress (Table 3). Specificinfluence of texture was investigated by examining themean porosity and porosity frequency distribution ofdifferent textures at different depth ranges.

Figure 7a plots the means of porosities of differentlimestone textures as a function of depth. At shallow anddeep depths, average porosity of the different textures isquite similar. At intermediate depths (1.5–2.5 km;5500–8500 ft), different limestone textures at the samedepth range have different mean porosities. Grain-stones appear to have consistently higher mean poros-ity than packstones, wackestones, and mudstones overthis interval. These porosity differences between pack-stones, wackestones, and mudstones are not consistentfrom depth interval to depth interval, so no systematicpattern is evident from the data. This inconsistencyprobably results from the distinctly non-normal distri-bution of porosity within each texture type (Figure 8).

The fraction of high-porosity samples (porosity >8%)in different limestone textures has systematic differencesthat are not evident from the analysis of mean porosity(Figure 7b). At depths <1.5 km (5000 ft), all limestonetextures have high percentages of samples with highporosity. From 1.5 to 2.6 km (5000–8500 ft), the grain-stones and packstones have a greater percentage of high-porosity rocks than do the mudstones and wackestones.Below 2.6 km (8500 ft), all limestone textures have a lowpercentage of samples with porosity >8%. The system-atic differences in economic porosity over the intermedi-ate depth range are especially significant because nosystematic trend in mean porosity of the different lime-stone textures was evident over the same depth interval.

The preservation of higher porosities in the grain-stone and packstone textures relative to the mud-supported textures, without significant differences in themean porosity, indicates that the distributions of porosi-ties within the various fabric types are strikingly differ-ent. Grain-supported rocks have a tail of high-porosityvalues that is missing or much smaller in the mud-supported rocks in samples >2.1 km (7000 ft) (Figure 8).

Effect of Temperature and Effective Stress

The average porosity in clean limestones appears todecrease with increasing thermal gradients at almostall depths (Figure 9). In the general linear model, tem-perature is of marginal statistical significance for themodel of limestone porosity as a function of depth,fabric, well, and temperature (Table 3). Although theoverall effect of temperature is small, in the depthrange between 2.6 and 3.0 km (8500–10,000 ft), thetemperature effect is significant. The effect is believedto be most significant in this range because at shal-lower depths, temperature differences are very smallfor the different thermal gradient ranges; deeper in thebasin, the average porosity in all wells is low.

Pore pressure variations are not significant com-pared to the geostatic load, so effective stress anddepth are highly correlated. Linear regression between

Figure 3. Mean porosity of carbonate lithologies as afunction of present-day burial depth. Symbols rep-resent average depth and average porosity for thatdepth and lithology. ARG. LS. = argillaceous lime-stone; LS. = clean limestone; DOL. LS. = dolomiticlimestone; DOL. = dolomite.

Page 54: Reservoir Quality Prediction in Sand and Carbonates

clean limestone porosity and the three depth-relatedvariables (depth, effective stress, and temperature)yields very similar R2 values. Given the strong covari-ance between these variables, it cannot be demon-strated unambiguously that effective stress is theprimary control on mean porosity rather than someother factor correlated with present-day burial, such astemperature or stress history.

DISCUSSION

Controls on Porosity

The data generated in this study demonstrate thatMississippian carbonates of the Williston Basin exhibitsystematic porosity loss with increasing depth of bur-ial. Although secondary porosity may increase poros-ity locally, or other secondary processes may result inincreased permeability, which increases the economicpotential of a carbonate, the overall trend is onetoward decreasing average porosity with increasingdepth of burial for all carbonate rock types. Effectivestress, temperature, stress history, or some other vari-able correlated with depth actually causes the reduc-tion in porosity, not the depth itself. Because of thestrong covariance of these variables in this data set, theactual relative influence of the different variables can-not be ascertained.

The difference between the basin-scale porositytrend and the reservoir porosity trend is quite striking.The porosity reduction with depth in economic lime-stone petroleum accumulations in the Madison Groupis nowhere near as great as the average porosity reduc-tion in all Madison Group limestones (Figure 10).Instead, average porosity of petroleum accumulations

in limestone has a modest systematic decrease in aver-age reservoir porosity with increasing depth.Dolomitic limestones and dolomites have averagereservoir porosity higher than that of limestones in thedeeper part of the basin.

The major control on rate of average porosity losswith depth is carbonate mineralogy. Dolomite anddolomitic limestones lose porosity at a slower rate thando limestones (Figure 4). Because of the location of con-trol wells, the sample size of dolomite lithology is toosmall for quantitative evaluation; however, the decreas-ing average porosity from dolomite through dolomiticlimestone to limestone indicates that the dolomiteaverage porosity estimates are consistent with porositytrends in other data. The progressive decrease in poros-ity with decreasing dolomite content over a narrowdepth range was also noted by Kupecz (1984, her figure45) in Billings anticline fields.

Argillaceous limestones have a lower overall poros-ity and a faster rate of porosity loss than do the cleancarbonates at similar depths (Figure 3). Porositybecomes lower as the clay content of the limestoneincreases (Figure 6; Table 2).

Porosity differences between rock types appears to bepredominantly a burial feature, not inherited from initial(zero-depth) porosity. Extrapolation of the exponentialporosity trends to zero burial depth gives an estimate ofzero-depth porosity. Zero-depth porosities of argilla-ceous limestones are highest, whereas dolomites havethe lowest zero-depth porosity. This is the oppositeporosity ranking than that seen over the entire depthrange of study, but this ranking is consistent with someobservations of the Cenozoic average porosity trends.

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 37

Figure 4. Porosity trends on asemilogarithmic plot. Theporosity loss plots close to astraight line in all of thelithologies, indicating thatthe porosity loss functionsare approximately exponen-tial. Low-porosity averageshave a large scatter on semi-logarithmic plots due to theinaccuracy of log-determinedporosity at low values. ARG.LS. = argillacious limestone;LS. = clean limestone; DOL.LS. = dolomitic limestone;DOL. = dolomite.

Page 55: Reservoir Quality Prediction in Sand and Carbonates

38 Brown

For example, average shallow-buried dolomites have alower porosity than coexisting limestones (Halley andSchmoker, 1983). If the zero-depth porosity is an indica-tor of preburial porosity, then the preburial porosity ofthe different carbonate lithologies is not responsible fortheir relative porosity in the subsurface.

In contrast to the mineralogy effect, texture has littlesystematic effect on average limestone porosity. At shal-low depths of burial, all limestone textures have similaraverage porosity (Figure 7a). At intermediate depths,texture has an effect on average porosity, but it is notsystematic. This has two implications: (1) differences in

Figure 5. Plots of porosity with depth for the carbonate lithologies, showing mean porosity (dots), ±1 standarddeviation (horizontal bars), and maximum porosity (triangles). The number at each depth refers to the numberof measurements of that lithology in that depth interval. (a) Limestone, (b) argillaceous limestone, (c) dolomiticlimestone, and (d) dolomite. Dashed lines are porosity trends of the dolomite and dolomitic limestone data.

Page 56: Reservoir Quality Prediction in Sand and Carbonates

average porosity in different limestone textures developduring burial diagenesis; they are not exclusively fea-tures inherited from initial porosity or early diagenesis;and (2) the erratic effect of texture on porosity may indi-cate that the textural features do control porosity loss,but the subdivisions used here are too broad to identifythe specific depositional fabrics and facies that selec-tively preserve porosity.

However, the fraction of rock with high porosity isclearly a function of texture (Figure 7b). Moderately todeeply buried grain-supported rocks are much morelikely to have high porosity than are mud-supportedrocks buried to the same depth; that is, a deeply buriedlime packstone is likely to have an average porosityquite similar to that of a lime wackestone or lime mud-stones buried to the same depth (Figure 7a). However,some beds in the lime packstones will have highporosity, whereas none of the beds in the lime wacke-stones and lime mudstones will.

One of the major results of this study is to documentthat average porosity in grain-supported and mud-supported limestones is about the same at equivalentburial depths. Only the fraction of grain-supported lime-stones that retain substantially higher-than-averageporosity can become economic limestone reservoirswithout fracturing. Porosity of some grain-supportedrocks is not occluded as rapidly as the porosity of othergrain-supported rocks or mud-supported rocks. As aresult, the near-normal distribution of porosity charac-teristic of shallow-buried, grain-supported limestonesevolves into a negatively skewed distribution with bur-ial (Figure 8). This process is treated statistically here,but deterministic, physicochemical processes such as

timing of mineralogical stabilization, early cementation,early diagenetic fabric alteration, or pore size or geome-try actually control which grain-supported rock maypreserve its porosity with burial.

Mechanisms of Porosity Loss

The porosity trends indicate that porosity-destructiveprocesses dominate over burial secondary porosity cre-ation. As in other shelf limestones, the Madison Groupcarbonates clearly lose porosity by cementation ratherthan mechanical compaction. This is evident from thin-section photomicrographs of numerous diagenetic andfield studies in the basin. The most likely process gener-ating the cement is pressure dissolution near the site ofcementation. This interpretation is supported by the cor-relation of porosity loss with clay mineral content. Clayminerals have been postulated to increase the effective-ness of pressure dissolution (Weyl, 1959).

Allochthonous carbonate cementation (i.e., carbon-ate cements derived from parts of the basin removedfrom the site of cementation by a distance of kilome-ters or hundreds of meters of burial) cannot be ruledout, but it is judged unlikely for two reasons. First,porosity loss is correlated with depth and not geo-graphic position. If allochthonous cements were pre-cipitated from moving water, there should be anasymmetry of cement distribution related to positionof recharge and discharge of the water, and patternsshould be less dependent on depth. Second, eachlithology follows its own porosity-reduction pathway.If reduced by allochthonous cement, all lithologiesshould follow a similar porosity-reduction path, so

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 39

Figure 6. Plot of density porosity against shaliness as indicated by gamma-ray (GR) intensity. (a) Shallow car-bonates (ARCO Wunderlich #1) show porosity decreasing with increasing shaliness as indicated by GR inten-sity. (b) Porosity in the deeply buried Amoco Sondrol #1 well has no systematic relationship to shaliness asindicated by the GR intensity. Lack of correlation is due to the small overall porosity variation in this argilla-ceous carbonate.

Page 57: Reservoir Quality Prediction in Sand and Carbonates

40 Brown

that the main influence is the initial porosity or pore-size distribution, not the mineralogy of the host rock.

Other results of this study provide insights into theproperties of the porosity-reducing process, whateverthe process turns out to be. The fact that averageporosity loss follows an approximately exponentialdecrease is evidence that the process of porosity losshas some sort of feedback. It is possible that the rate ofporosity loss is proportional to the porosity of therock. It is also possible that increasing depth or tem-perature actually decreases the rate of porosity loss inlimestone by increasing rock ductility (decreasingstress differences, which may lead to dissolution) orby locking of stylolite surfaces and decreasing the areaat which cement is generated.

Dolomites are generally found to lose porosity withdepth at a slower rate than limestones of the same agein this basin (Figure 3). This clearly indicates thatdolomite porosity is being selectively preserved withrespect to limestone. Dolomite porosity in Paleozoicbasins higher than porosity of surrounding limestonehas been interpreted as evidence that dolomitizationcreates porosity (Weyl, 1960). Whether dolomitizationactually creates porosity in geological settings remainscontroversial, but the porosity vs. depth trends for thisbasin substantiate the selective preservation mecha-nism as the dominant reason for the difference in lime-stone and dolomite porosity, at least in this Paleozoicformation. It is postulated that the lower rate of porosityloss in dolomite is related to its higher bulk modulus(Ellis et al., 1988), or related to slower diffusion or pre-cipitation kinetics. A higher bulk modulus results indecreased chemical potential change with stress duringpressure dissolution (Paterson, 1973).

Porosity Prediction from Average Porosity and Porosity Distribution

The type of data presented here can be used to predictthe average porosity and the proportion of high (i.e.,reservoir) porosity at a given depth. Such predictionscan be used for two purposes: (1) estimation of velocities

and densities as a function of depth for synthetic seismicreflection data sections and (2) estimation of both poros-ity risk and reserves for prospects with a given depth,lithology, and fabric.

Average porosity can be used to estimate averagerock properties (density, velocities) where no site-specific data (sonic logs or vertical seismic reflectiondata profiling) are available. The porosity range can beused to estimate the range of acoustic impedance pos-sible within a single lithology at a depth. Ranges ofporosity at a given depth can also be used to model anexpected range of acoustic impedance at boundariesbetween two carbonate lithologies.

Although average porosity is a qualitative indicatorof the likelihood of finding economic porosity, theaverage porosity is not likely to be the average of thepetroleum-charged porosity. Petroleum selectivelycharges the fraction of the rock with the best-qualityrock properties within the trap. Also, those sections oflow reservoir quality that are charged with petroleummay not be economically recoverable. For these rea-sons, economic porosity in many settings (especiallyPaleozoic carbonates) is usually substantially greaterthan the average porosity.

The fraction of the total rock likely to exceed an eco-nomic threshold porosity is a better predictor for thelikelihood and amount of economic porosity, becausethis is the fraction of rock that is likely to be charged bypetroleum if trapping conditions are favorable. If adistribution type (such as normal, log normal, orCauchy) is assumed, it can be fitted to the frequencydiagrams for porosity at different depths. Theexpected fraction of the carbonate section exceedingan arbitrarily chosen porosity threshold can then bedetermined for the distribution (Figure 11).

A simple, empirical method of estimating the frac-tion of economic porosity in an interval with a givenfabric or lithology is proposed as an alternative proce-dure that requires no assumption of a distribution.Porosity data are plotted on a cumulative frequencyplot, with frequency plotted on a probability scale, and

Table 3. Effects of Variables on Limestone Porosity.

Variable Sum of Squares Significance*

Model 1: Limestone Porosity = φ (Depth, Fabric, Effective Stress, Temperature)**

Mode 8747.9 0.0001 (R2 = 0.5)Error 8744.6 ---Depth (ft) 5.3 3 0.426Fabric (class variable) 98.03† 0.02Effective Stress (psi) 148.16† 0.0001Temperature (°F) 10.7† 0.2702

Model: Limestone Porosity = φ (Well, Fabric, Effective Stress, Depth, Temperature)**

Mode 11,142.8 0.0001 (R2 = 0.637) Error 6349.6 ---Well (class variable) 10,380.9 (2395†) 0.0001

*Probability of the regression coefficient to equal zero as determined from the t-test; the lower thenumber, the greater the significance of the variable. **Regression coefficients cannot be shown here because GLM model is a type of ANOVA. The purpose of table is to show signficance of variables. †Sum of squares if added last to the model in stepwise fashion.

Page 58: Reservoir Quality Prediction in Sand and Carbonates

the porosity plotted on either a linear or a logarithmicscale. The economic threshold porosity is chosen onthe horizontal axis, based on drilling experience in thearea. The cumulative porosity greater than the thresh-old is determined from the porosity cumulative fre-quency distribution by reading the cumulativeprobability on the vertical axis (Figure 11).

As an example, the porosity cumulative frequencydistributions for mudstones and packstones at differentaverage depths are plotted on a linear porosity scale(Figure 12a). Due to the small sample size of porosities ofgiven depths and lithologies, considerable scatter occursnear the tails of the distributions, so smoothed curveswere fitted to the tails of the data by sight (Figure 12b). Ifdesired, more quantitative fits could be made with anassumed distribution. For our example, let us assumethat a minimum of 8% porosity is needed in an oil-saturated reservoir at 3 km (10,000 ft) to be consideredpay. Figure 12b indicates that the fraction of mudstonesexceeding this porosity is off the scale, and estimated tobe about 0.01%. In contrast, packstones should haveabout 3% of the section with porosity exceeding 8%. If 30 m (100 ft) of each lithology were penetrated, 3 mm(~0.1 in.) of the mudstone interval would have porositygreater than 8%, whereas 1 m (~3 ft) of the packstonewould exceed this porosity. If a minimum of 3 m (10 ft)of pay were required for economic production, thechances of encountering adequate thickness and

porosity in the mudstone would be remote. Althoughthe risk for encountering the same thickness of porosityin the packstone may be high, it may be acceptable if eco-nomic factors are favorable, or if an especially thick sec-tion of packstone were expected from facies models.

Several points should be made about the previousexample and this approach to reservoir quality predic-tion. First, it is essential that the porosity distributionbe developed from unbiased data. If, for example,porosity data were collected only from intervals withsome minimum threshold porosity, then the fractionof economic porosity is relative to the total thickness ofthe porosity with the minimum porosity level, not rel-ative to the thickness of the carbonate body as a whole.This gives the prediction much greater uncertainty,because the fraction of the total carbonate thicknesswith the threshold minimum porosity level must alsobe estimated. Likewise, if only maximum porositydata are collected, porosity prediction has less power,because much production from economic reservoirscomes from intervals below the maximum porosity.

Second, porosity in carbonates is typically spatiallycorrelated in vertical sections; that is, high-porosity sam-ples tend to lie near other high-porosity intervals in avertical section. The problem is that the thickness of thecorrelated interval (referred to as “bed” here) is notknown. In the previous example (Figure 12), 1 m of eco-nomic porosity was predicted in the packstone interval.

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 41

Figure 7. Porosity of limestone fabrics. (a) Mean porosity of fabrics in depth intervals. There is no consistentranking of the mudstone, wackestone, or packstone porosities from depth interval to depth interval. Thegrainstone samples are too small to be statistically distinguished from the packstones at the same depth giventhe large standard deviations of the porosity means. (b) Percentage of high-porosity samples (φ >8%) in car-bonate fabrics from different depth intervals. Between 7.5 and 2.5 km, (500–8500 ft), the framework-supportedcarbonates have a significantly higher percentage of samples with porosities >8%. Below 2.5 km (8500 ft), allcarbonates have low percentages of high-porosity samples

Page 59: Reservoir Quality Prediction in Sand and Carbonates

42 Brown

It is possible that this occurs as five beds each 20 cmthick, as a 1-m-thick bed, or as some other distribution.The analysis presented here will not distinguishbetween the alternatives. Statistical analyses of verticalsections in both carbonate and siliciclastic intervals indicate that as the interval between samples decreases,the porosities of the samples are more likely to be corre-lated (Kittridge et al., 1990). Results of these types ofstudies indicate that vertical spatial correlation is on theorder of feet to a few tens of feet (1–10 m) (Kittridge etal., 1990).

Vertical spatial correlation of porosity varies fromlocation to location, and can best be calibrated by localdata. If an average spatial correlation is assumed fromlocal calibration data, the risk of finding a thickness ofeconomic porosity greater than the expected economicporosity thickness can be calculated by the binomialsampling theory. The number of independent trials isthe gross thickness of the unit divided by the thicknessof independent units as estimated from semivari-ograms, and the probability of success of each trial isdetermined from the porosity cumulative frequencydistribution, as discussed above.

Williston Basin Porosity Prediction

The data set collected here can strictly be used onlyfor prediction of porosity in Mississippian carbonatesof the Williston Basin. The preferred method for wild-cat exploration is described in the previous section.From the calculations presented here, the depth limitfor economic (>8%) porosity in grain-supported Mis-sissippian limestones of the Williston Basin is ~2.5 km(8000 ft; 20% probability) to 3 km (10,000 ft; 2% proba-bility), depending on which probabilities are at accept-able risk for random drilling.

Although specific compositional or textural subdi-visions have been able to distinguish depositionalfacies with significantly higher porosity in field stud-ies (e.g., pisolitic facies at Glenburn field; Gerhard,1985), the gross textural subdivisions used here could

not. In many exploration settings, these sorts of grosstextural subdivisions are likely to be the only availableinformation. This means that one rarely has the higherquality depositional facies information necessary topredict the presence of a depositional facies shown tohave higher average porosity.

Williston Basin Madison Group fields withdolomite reservoir rocks are concentrated in the south-ern and western part of the basin. For this reason, thebasinwide sampling pattern resulted in too fewdolomite samples to apply the cumulative frequencyanalysis approach used for the limestones. The highervariability of the dolomite average porosity vs. depthis interpreted to be caused by small sample size and amore variable diagenetic history than the limestone.Porosity does not seem to be a problem with MadisonGroup dolomites down to the maximum depth exam-ined as part of this study.

Although strictly applicable to the Williston BasinMississippian rocks, the trends developed from thesedata can also be used as a guide to porosity predictionin other Paleozoic cratonic basins. Specifically, a highrisk for limestone porosity is expected in Late Paleozoicreservoirs buried much deeper than 3 km (10,000 ft).Late Paleozoic dolomite reservoirs are not expected tohave much of a reservoir quality problem due to burialcementation down to 3 km. However, the significanceof evaporitic cementation on porosity was not evalu-ated in this study, and it is likely that anhydrite orhalite cementation could significantly reduce porosityfor those dolomites associated with evaporitic sections.Of course, dolomites buried with low initial porosityare not likely to develop substantial porosity with bur-ial, so these results can only be applied to dolomiteswith high initial porosity.

Comparison with Other Porosity Trends

Comparison of different chalk and limestone porosity–depth trends and porosity–TTI trends indicates thatporosity data from one basin cannot be directly used to

Figure 8. Porosity distribu-tion histograms of samplesfrom representative depths.The histograms indicate thenumber of samples (verticalscale) with the given porosi-ty (horizontal scale, in per-centages). The grainstonesamples are indicated bycross hatch in the packstoneand grainstone histograms.

Page 60: Reservoir Quality Prediction in Sand and Carbonates

estimate carbonate porosity in another basin (Schmoker,1984). However, where basin-scale, shallow-water lime-stone data are compared between basins, older lime-stones have an average porosity that is lower than thatof younger limestones at the same maximum burialdepth (Figure 13). This indicates that time is importantfor porosity reduction, in addition to effective stress and

temperature, as postulated by Schmoker (1984). Theeffect appears to be somewhat systematic, and provideshope that a generic fundamental relationship betweenlimestone porosity and burial can be developed.

Although the relative magnitudes of the effects oneffective stress, time, and temperature on porosity losscannot be ascertained from this study due to its

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 43

Figure 9. Temperature gradient effect on limestoneporosity. Gradients are divided into high (circles),medium (squares), and low (triangles), as discussedin the text.

Figure 10. Comparison of average limestone porosi-ty trend developed here (solid line) with reservoirporosity of Madison Group fields of different reser-voir mineralogy. Some fields with limestone may bedolomitic limestone or dolomite.

Figure 11. Cumulative frequencydiagram with normal, log normal,and empirical distributions, show-ing different estimates of porositygreater than a threshold porosityfor different distributions with thesame median porosity and similarstandard deviation. Because mostporosity sample sets are not suffi-ciently large to use as a direct cor-relation for very high porosityvalues, the cumulative frequencytrend has to be extrapolated wherehigh-porosity samples constitute asmall fraction of the total porositypopulation.

Page 61: Reservoir Quality Prediction in Sand and Carbonates

44 Brown

Figure 12. Sample porosity cumulative frequency distributions for packstones (a) and mudstones/wackestones(b) plotted on a probability scale. Smoothed porosity cumulative frequency distributions for packstones (c)and mudstones/wackestones (d). Numbers along the cumulative frequency curves correspond to depth range:4 = 900–1200 m (3000–4000 ft); 5 = 1200–1500 m (4000–5000 ft); 6 = 1500–1800 m (5000–6000 ft); 7 = 1800–2100 m(6000–7000 ft); 8 = 2100–2400 m (7000–8000 ft); 9 = 2400–2700 m (8000–9000 ft); 10 = 2700–3000 m (9000–10,000 ft);and 11 = 3000–3300 m (10,000–11,000 ft).

Page 62: Reservoir Quality Prediction in Sand and Carbonates

design, it has been demonstrated that temperaturedoes have an effect on porosity loss independent oftime and effective stress. Temperature and time havebeen postulated to be the major controls (Schmoker,1984), but this cannot be verified in this study, andregression equations seem to indicate that effectivestress is still dominant over temperature.

CONCLUSIONS

Average porosity decreases as a function of depthin Mississippian carbonates from the Williston Basin.This porosity decrease is approximately exponentialfor all carbonate lithologies. The greatest control onrate of porosity loss with depth is the lithology of thecarbonate. Argillaceous limestones lose porosity atthe greatest rate and have the lowest porosity at alldepths analyzed. Clay-free limestone porositydecreases faster with depth than does dolomitic lime-stone porosity, and dolomite porosity decreases theleast with burial depth. The effect of limestone fabricon average porosity is quite small, but fabric has astrong influence on range of porosity at a given depth,and thus on the presence of high (economic) porosity.In Williston Basin Mississippian limestones, the selec-tive occurrence of economic porosity in grain-supported rocks is due to the selective preservation of

porosity in a small fraction of the grain-supportedrocks, while porosity in most grain-supported rocksand all mud-supported rocks is systematicallydestroyed. The exact geological mechanism for selec-tive preservation of porosity cannot be determinedfrom this type of study. Increased thermal gradientenhances porosity loss in limestone.

The expected net thickness of economic porosity canbe estimated from cumulative frequency distributionsof porosity samples. The distributions are skewed sig-nificantly in moderately to deeply buried samples, so anormal distribution cannot be assumed for predictionof abundance of high porosity. If the average thicknessof beds with similar porosity levels can be estimated,these estimates can be converted into quantitative riskfactors using standard binomial sampling theory.

Because the Williston Basin is a well-drilledpetroleum province, the application of this study tothe Williston Basin is limited. In most parts of thebasin, porosity can be mapped and the drilling loca-tion chosen to enhance the likelihood of encounteringadequate porosity. The drilling is not random, so theodds of encountering porosity are significantly greaterthan those presented here, assuming reservoir qualityof nearby wells is carefully assessed. However, thisstudy demonstrates the method by which porosityloss in carbonates in other, less well drilled settingscan be evaluated and the method by which scatter ofporosity data can be used to predict the risk forencountering porosity exceeding a threshold value.

The results seen here confirm the general trendsobserved elsewhere. (1) Average carbonate porositydoes decrease with depth (Schmoker and Halley,1982). (2) Limestones lose porosity with depth at afaster rate than do dolomites with equivalent burialhistories (Schmoker and Halley, 1982). (3) Averageporosity of a limestone at a given depth decreases withincreasing age (Schmoker, 1984). These generaliza-tions can be used as a guide to evaluate new deepplays for which little empirical data are available.

ACKNOWLEDGMENTS

The author thanks Bob Loucks, Jim Hickey, JulieKupecz, James Schmoker, Jerry Lucia, and AndrewHorbury for reviews. I also thank ARCO Explorationand Production Technology Co. for permission torelease this study, which was completed as an internalstudy in 1984. Gulf Coast Cretaceous porosity vs.depth data were provided by Bob Loucks.

REFERENCES CITED

Beach, D.K., and J.W. Griffin, 1992, Stanley field—U.S.A.(Williston Basin, North Dakota), in N.H. Foster andE.A. Beaumont, compilers, Stratigraphic traps III:Tulsa, Oklahoma, AAPG Treatise of Petroleum Geol-ogy, Atlas of Oil and Gas Fields , p. 389–420.

Choquette, P.W., and L. Pray, 1970, Geologic nomen-clature and classification of porosity in sedimentarycarbonates: AAPG Bulletin, v. 54, p. 207–250.

Porosity Variation in Cabonates as a Function of Depth: Mississippian Madison Group, Williston Basin 45

Figure 13. Limestone porosity trend compared to otherquantitative porosity trends for shelf limestones.Florida data from Schmoker (1984). Texas Cretaceousdata are unpublished core analysis trends collected byR.G. Loucks (1985, personal communication).Ordovician data are average wireline-log limestoneporosity from three wells penetrating the Red RiverFormation, collected as part of this study.

Page 63: Reservoir Quality Prediction in Sand and Carbonates

46 Brown

DeFord, R.K., et al., eds., 1976, Geothermal gradient mapof North America: Tulsa, Oklahoma, AAPG, scale1:5,000,000, 2 sheets.

DeMis, W.D., 1992, Elkhorn Ranch field—U.S.A. (Willis-ton Basin, North Dakota), in N. H. Foster and E. A.Beaumont, compilers, Stratigraphic traps III: Tulsa,Oklahoma, AAPG Treatise of Petroleum Geology,Atlas of Oil and Gas Fields, p. 369–388.

Dunham, R.J., 1962, Classification of carbonate rocksaccording to depositional texture, in W. Ham, ed.,Classification of carbonate rocks: AAPG Memoir 1,p. 108–121.

Ellis, D., J. Howard, C. Flaum, D. McKeon, H. Scott, O.Serra, and G. Simmons, 1988, Mineral logging para-meters: nuclear and acoustic: Technical Review, v. 36,p. 38–52.

Gerhard, L.C., 1985, Porosity development in the Missis-sippian pisolitic limestones of the Mission CanyonFormation, Glenburn field, Williston Basin, NorthDakota, in P.O. Roehl and P.W. Choquette, eds., Car-bonate petroleum reservoirs: New York, SpringerVerlag, p. 192–205.

Halley, R.B., and J.W. Schmoker, 1983, High-porosityCenozoic carbonate rocks of south Florida: progres-sive loss of porosity with depth: AAPG Bulletin, v. 67,p. 191–200.

Kittridge, M.G., L.W. Lake, F.J. Lucia, and G.E. Fogg,1990, Outcrop/subsurface comparisons of hetero-geneity in the San Andres Formation: SPE FormationEvaluation, September 1990, p. 233–240.

Kupecz, J., 1984, Depositional environments, diage-netic history, and petroleum entrapment in the Mis-sissippian Frobisher-Alida interval, Billingsanticline, North Dakota: Colorado School of MinesQuarterly, v. 79, no. 3, 62 p.

LeFever, R.D., and J.A. LeFever, 1991, Newburg andSouth Westhope fields—U.S.A. (Williston Basin,North Dakota), in N.H. Foster and E.A. Beaumont,compilers, Stratigraphic traps II: Tulsa, Oklahoma,AAPG Treatise of Petroleum Geology, Atlas of Oiland Gas Fields, p. 161–187.

Lindsay, R.F., 1985, Madison Group (Mississippian)reservoir facies of Williston Basin, North Dakota:AAPG Bulletin, v. 69, p. 279–280.

Lindsay, R.F., and C.G.St.C. Kendall, 1985, Deposi-tional facies, diagenesis and reservoir characterof Mississippian cyclic carbonates in the MissionCanyon Formation, Little Knife field, Williston

Basin, North Dakota, in P.O. Roehl and P.W.Choquette, eds., Carbonate petroleum reservoirs:New York, Springer-Verlag, p. 177–190.

Miller, W.R., and S.A. Strauz, 1980, Preliminary mapshowing freshwater heads for the Mission Canyonand Lodgepole Limestones and equivalent rocks ofMississippian age in the Northern Great Plains ofMontana, North Dakota, South Dakota, andWyoming: U.S. Geological Survey Open File Report80–729, map, 1 sheet.

Paterson, M.S., 1973, Nonhydrostatic thermodynamicsand its geologic applications: Reviews of Geo-physics and Space Physics, v. 11, p. 355–389.

Peterson, J.A., and L.M. MacCary, 1987, Regional stratig-raphy and general petroleum geology of the U.S. por-tion of the Williston Basin and adjacent areas, inWilliston Basin, in M.W. Longman, ed., Anatomy of acratonic oil province: Denver, Colorado, RockyMountain Association of Geologists, p. 9–44.

Schmoker, J.W., 1984, Empirical relation between car-bonate porosity and thermal maturity: an approachto regional porosity prediction: AAPG Bulletin, v. 68,p. 1697–1703.

Schmoker, J.W., and R.B. Halley, 1982, Carbonateporosity vs. depth: a predictable relation for SouthFlorida: AAPG Bulletin, v. 66, p. 2561–2570.

Schmoker, J.W., K. Krystinik, and R. Halley, 1985,Selected characteristics of limestone and dolomitereservoirs in the United States: AAPG Bulletin, v. 69,p. 733–741.

Scholle, P.A., 1978, Porosity prediction in shallow vs.deep water limestones: 53d Annual Fall TechnicalConference of the Society of Petroleum Engineers,Houston, Texas, October 1978, SPE Preprint SPE7554, 6 p.

Tonnesen, J.J., 1985, ed., Montana oil and gas fields:proceedings (2 volumes): Billings, Montana, Mon-tana Geological Society, 1217 p.

Tyler, C.D., ed., 1962, Oil and gas fields, North DakotaSymposium: Bismarck, North Dakota, NorthDakota Geological Society, 220 p.

Weyl, P.K., 1959, Pressure solution and the force ofcrystallization—a phenomenological theory: Journalof Geophysical Research, v. 64, p. 2001–2025.

Weyl, P.K., 1960, Porosity through dolomitization:conservation of mass requirements: Journal of Sedi-mentary Petrology, v. 30, p. 85–90.

Page 64: Reservoir Quality Prediction in Sand and Carbonates

47

Chapter 4

Predicting Reservoir Quality Using LinearRegression Models and Neural Networks

K.M. LoveExxon Production Research Co.

Houston, Texas, U.S.A.

C. StrohmengerBEB Erdgas und Erdöl GmbH

Hannover, Germany

A. WoronowExxon Production Research Co.

Houston, Texas, U.S.A.

K. RockenbauchBEB Erdgas und Erdöl GmbH

Hannover, Germany

ABSTRACT

A method for predicting the three-dimensional distribution of reservoirattributes has been developed by integrating geological and statistical mod-els. The general method, applicable to carbonate and siliciclastic reservoirs,has been demonstrated by predicting the distribution of dolomite, calcitizeddolomite, porosity, and permeability from regional to field scales in thePermian Zechstein 2 Carbonate of northern Germany.

The first step in the prediction process consists of identifying factorspotentially responsible for reservoir quality distribution. For the Zechstein 2Carbonate, the resulting geologic model suggested that paleofaults andrelated fracture systems controlled the distribution of nonporous calcite (cal-citized dolomite) by acting as conduits for calcitizing fluids originating fromanhydrites underlying the carbonates.

The next step in the prediction process involves determining if the geologicmodel provides variables that can be used to predict the variable of interestgiven the predrill data available. If not, then other predictor variables, not nec-essarily cause-and-effect variables but ones whose values are known predrill,are required. Although a geologic model for Zechstein diagenesis elucidatedthe probable cause-and-effect relationship regarding the distribution of

Love, K.M., C. Strohmenger, A. Woronow, and K.Rockenbauch, 1997, Predicting reservoir qualityusing linear regression models and neural net-works, in J.A. Kupecz, J. Gluyas, and S. Bloch,eds., Reservoir quality prediction in sandstonesand carbonates: AAPG Memoir 69, p. 47–60.

Present address: Exxon Exploration Co., Houston, Texas, U.S.A.

Page 65: Reservoir Quality Prediction in Sand and Carbonates

48 Love et al.

INTRODUCTION

Geologic studies commonly provide a means to linkreservoir quality to one or more controlling factors. Ifthese factors subsequently predict predrill reservoirquality at the necessary scale, reservoir risk can bereduced. In some cases, however, the controlling fac-tors cannot be identified, or, more commonly, knowl-edge of the controlling factors does not permitprediction at a pragmatic scale. For such cases, quanti-tative models based on linear or nonparametric meth-ods that rely, at least in part, on location variables (x-ycoordinates and depth) may provide a useful meansfor predicting the three-dimensional distribution ofnonrandomly distributed parameters.

This chapter gives a general approach to the resolu-tion of such pragmatic prediction issues, using a casestudy for illustrative purposes. The Upper PermianZechstein 2 Carbonate of northern Germany (Figure1) provides an example of a reservoir-quality problemwhere cause-and-effect models failed to generate

practical predictions. Carbonates of the second Zech-stein cycle (the Ca2 or Stassfurt Carbonate) constitutenorthern Germany’s most prolific carbonate gas play;consequently, many efforts focus upon characterizingreservoir quality. Prediction of depositional faciesprovides one key to reservoir-quality prediction(Strohmenger et al., 1996), but the Zechstein 2 Carbon-ate underwent an extensive calcitization of dolomite(“dedolomitization”) (Figure 2) that generallydestroyed porosity and permeability and was notdepositional-facies-specific (Strohmenger et al., 1993).Calcitization generally increases basinward, but itslateral distribution has been difficult to predict withinindividual slope facies. Thus, predicting the distribu-tion of this nonporous diagenetic calcite vs. porousdolomite was identified as a crucial first step towardpredicting reservoir quality, especially within thickslope deposits. As a result, a geologic model wasdeveloped to explain the mechanism of calcitizationin the hope of using the model to predict calcite vs.dolomite. The resulting model indicated, however,

mineral types, it provided no means for predicting the geographic distributionof mineral types, because data on the distribution of paleofault and paleofrac-ture systems cannot be obtained. For pragmatic purposes, models must bothpredict the desired parameter at the necessary scale and use predictor vari-ables whose values are known prior to drilling. For the Zechstein 2 Carbonate,linear regression models using facies and location (x-y coordinates and depth)accomplished practical predictions of mineral distribution. The fact that loca-tion provides significant predictions indicates that calcite and dolomite occurin a spatially organized manner, reflecting the geologic processes that causedthe calcitization of the dolomite. Because paleostructure presumably con-trolled calcite distribution, separate models were developed for structurallydistinct subareas. The use of structural subdivisions provided a way toaccount for different types of calcite distribution caused by different types offault and fracture systems.

Although mineralogy is a dominant control on reservoir quality in theZechstein 2 Carbonate, the porosity and permeability distributions reflectadditional factors. Like the mineralogy distribution, however, the porosity andpermeability distributions have a dominant nonrandom spatial component,and therefore can be predicted reliably using location information. Because thespatial distribution of porosity and permeability in the Zechstein 2 Carbonateis highly complex, a nonparametric predictive technique (an artificial neuralnetwork) was implemented. It produced models that surpassed those of linearregression.

Although cast here in terms of a particular application, the methodology isgeneral, and such predictive models can be used to generate maps and crosssections of predicted parameters within any reservoir. In addition, sets of pointvalues generated by the models can be loaded into visualization software toprovide three-dimensional representations of the predicted parameters.

Page 66: Reservoir Quality Prediction in Sand and Carbonates

that paleofaulting and paleofracturing were responsi-ble for the calcite distribution, thus providing no vari-ables that could be used directly to predict calcitedistribution at the desired scale prior to drilling. As aresult, statistical models were sought that used valuesof variables accessible before drilling as proxies forthe unattainable values of the cause-and-effect vari-ables. For the purpose of predicting calcite, linearregression models were used.

Although prediction of calcite improved reservoir-quality prediction in the case of the Zechstein 2 Carbon-ate, other factors influence porosity and permeabilitydistribution. Thus, models were developed to predictporosity and permeability distribution directly. Becauseof the functional complexity of the porosity and perme-ability distributions, artificial neural network models (aform of artificial intelligence) and linear regressionmodels were used. The objective of the models was topredict porosity and permeability in as much detail aspossible ahead of the drill. Although the models are notcapable of replicating the high-frequency variations ofporosity and permeability that occur within a facies,trends within facies can be predicted.

DATA

A statistical study of factors useful for predictingreservoir-quality distribution requires a database con-taining variables likely to be either directly or indirectlyrelated to reservoir quality. For the Zechstein 2 Carbon-ate study, an existing database at BEB Erdgas und ErdölGmbH was expanded to include data for hypothesizedreservoir-quality controls. Core data included mineral-ogy, facies, subfacies, porosity, and permeability from287 wells. The cores provided good coverage of thefacies present in a given area, and data from core plugsgenerally were available every 15 cm throughout a core.Although each core did not necessarily cover the entireCa2 interval, enough data from surrounding wells wereavailable to adequately represent all facies present in agiven area. Well log, structural, geochemical, thickness,and location data also were available. Because of thelarge amount of core available, all porosity and perme-ability values used for model development came frommeasurements on core plugs, rather than from welllogs. For reasons discussed later, the data were dividedinto ten subsets, ranging in number of wells from 7 to81, and in number of samples from 616 to 6990.

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 49

Figure 1. Location of studyarea (outlined in black)within the Upper PermianZechstein 2 Carbonate (Ca2)of the Southern ZechsteinBasin in northern Germany.LSW = lowstand wedge.

Page 67: Reservoir Quality Prediction in Sand and Carbonates

50 Love et al.

Data Smoothing

An important problem in predicting porosity andpermeability using core-plug data is the large varia-tion of the values over small distances within a core(high-frequency, high-excursion data). For example,porosity values commonly differ by an order of mag-nitude (e.g., from 2% to 20%) among several core plugsseparated by <1 m. These differences may be due togeologic factors (subfacies changes or diagenesis),sampling bias (choosing the unusual specimen foranalysis), or erroneous measurements.

No existing method has the ability to predict suchabrupt, centimeter-scale excursions, but prediction atthis level is generally not necessary. However, becausemodels cannot predict the abrupt change, the residualvalues (the observed porosity/permeability valueminus the model-predicted value) from these modelswill be large. To make the variance of the calibrationdata set more commensurate with the predicted data set,data can be “smoothed.” Kacewicz (1994) found thatsmoothing data improved the performance of a neuralnetwork. A simple way to accomplish data smoothinginvolves using an average value of porosity or perme-ability for each well or for each major subdivision withina well such as facies; however, models typically can pre-dict a higher frequency of variation than this, so infor-mation would be lost through such “oversmoothing.”

To avoid oversmoothing, a smaller window ofobservations for averaging can be used. In addition, amoving average can be calculated. Although thisdiminishes the abruptness of the changes in porosity-permeability values, a single high-excursion value stillhas a large influence. Thus, we used a tapered movingaverage. This weights the values closer to the middle ofthe window more heavily than those at the ends. Forthe Zechstein 2 Carbonate, a weighted moving averageof five measurements (a window typically <1 m) wascalculated so that the middle value was weighted mostheavily (0.4), then the two adjacent values less (0.2each), and the top and bottom values least (0.1 each).

The same procedure was used with the five depthsassociated with the five porosity and permeability val-ues to obtain an average depth value. The averagingprocedure did not cut across facies boundaries.

METHODS

Prediction Techniques

Although statistical procedures can use different vari-ables to predict the distribution of a parameter, the abilityto predict does not imply cause and effect. Establishingcause and effect is not necessary to accomplish the goalof prediction. All variables that significantly predictparameters of interest may aid in understanding causeand effect, but their utility in prediction models dependson the ability to estimate their values away from wellcontrol. If, for example, the percentage of anhydritecement in the Zechstein 2 Carbonate significantly pre-dicts carbonate mineralogy, this might give clues aboutthe calcitization process; however, cement contentwould be difficult to estimate in undrilled localities, soits input value in an equation to predict calcite vs.dolomite is uncertain, and the variable is not useful inpractical estimation. In fact, a poor estimate of the inputvalue may adversely affect predictive capabilities.

Regression

Regression analysis was used extensively to identifywhich variables significantly predict mineralogy, poros-ity, and permeability in the Zechstein 2 Carbonate. Inaddition, predictive models were developed using for-ward stepwise regression with backward elimination,which is a method to select a few predictor variablesfrom a large number of potential predictor variables.Using this procedure, individual variables enter intoand exit from an evolving model (Draper and Smith,1981; Bowerman and O’Connell, 1990). For the Zechsteinmodeling, a significance level of 0.15 was chosen for avariable to enter into and to remain in the model.

Figure 2. Schematic crosssection through theZechstein 2 Carbonateshowing distribution ofdepositional facies [plat-form, platform-LSW (low-stand wedge), upper slope,middle slope, lower slope,and basin] and mineralogy(dolomite vs. calcite). Thedistribution of calcite with-in the slope is not faciesdependent.

Page 68: Reservoir Quality Prediction in Sand and Carbonates

Although stepwise regression helps to reduce the num-ber of predictor variables, it does not necessarily providethe best regression model, nor can the remaining predic-tor variables be considered the most important.

The regression analyses made extensive use ofindicator variables [that is, variables that assume dis-crete values (e.g., 0 and 1)] to identify different cate-gories of a variable (Kleinbaum and Kupper, 1978). Inthis study, for example, an indicator variable was cre-ated for mineralogy, where mineralogy = 0 if the sam-ple (a core plug) is dolomite, but mineralogy = 1 if thesample is calcite. If regression models are constructedsuch that a 0,1 variable is the dependent (predicted)variable, and hypotheses about the regression will betested, it may be advisable to use a logistic function.However, if the purpose of the regressions is only to pre-dict, as is the case in this study, such transforms arenot mandatory.

Indicator variables allow the inclusion of such qual-itative data as mineral type and facies in quantitativemodels. The number of (0,1) indicator variablesrequired to represent one type of information (e.g.,facies) is n – 1, where n is the number of different cate-gories for the information (e.g., upper slope, platform,etc.) (Kleinbaum and Kupper, 1978). If five differentfacies occur, which is the case for the Zechstein 2 Car-bonate, then four indicator variables are required torepresent all the facies, as the “missing” variable isrepresented when the other four variables equal 0. Forthe Zechstein data, the lower slope facies was not usedas a facies variable; thus, when the platform, platform-lowstand-wedge, upper slope, and middle slope faciesvariables were all 0 for a particular sample, that sam-ple represented the lower slope facies.

If the relationship between the predicted and thepredictor variables is complex, continuous variables,such as depth or thickness, can be transformed to newvariables to possibly improve the regression model.Common transformations include logs, squares,square roots, and reciprocals of the original variables.Because of the complex spatial distribution of porosityand permeability in the Zechstein 2 carbonate, severaldifferent transformations of the location variableswere made in attempts to accommodate nonplanarvariations. For example, logarithms, squares, and reci-procals of the spatial data were offered to the predic-tion models, and commonly provided improvedpredictive ability (e.g., ln(depth), X2, 1/Y).

Artificial Neural Networks

Although prediction of calcite in the Zechstein 2Carbonate was relatively straightforward using linearmodels, prediction of porosity and permeability usinglinear regression was commonly improved by theaddition of terms higher than second order. This sug-gested a high level of complexity (nonlinearity?); forthis reason, an adaptive nonparametric predictionmethod was sought that might better predict the poros-ity/permeability distribution. One such method is anartificial neural network. A network consists of inter-connected computing cells; weights are assigned to theconnections between cells. These weights are used in

conjunction with input data to predict some outcome.The network learns to predict a desired outcome byiteratively modifying the weights and comparing thepredicted result with the actual result (Haykin, 1994).

This study used BPNET, a back-propagation neuralnetwork developed by author A. Woronow. The pro-gram randomly splits an input data file into a trainingdata set (from which the network learns how to predictporosity or permeability) and a test data set that is notused during training (to assess the prediction effective-ness of the network as learned from the training dataset). The test data consisted of ~10% of a data set. Aswith regression, the withheld data constitute a criticalpart of the evaluation of the predictive capabilities;they provide the only means to evaluate how well thepredictive tool will work when presented with newdata. The program was run on a 486-33 PC, and theneural network used one hidden layer, eight nodes,and a sigmoidal logistic function (Haykin, 1994).

For each data set, the network was allowed to learnuntil no further effective improvement occurred. Thetime required to reach this state ranged from ~30 min to3 hr, depending on the size and complexity of the dataset, although the program was allowed to run beyondthe cessation of improvement to ensure that a later“breakthrough” in learning did not occur. In two cases,the program was allowed to run for ~14 hr to furthercheck for this possibility; no additional learningoccurred. The normal training times corresponded tobetween 500 and 3000 learning cycles, where a cycle isone pass through each case in the training data set

Unlike for the regression models, formed variableswere not important for the neural network models,because a network can effectively develop its own linearand nonlinear transformations of variables to providebetter prediction. If, however, a known relationshipoccurs (e.g., it is known that one variable is related toanother by a particular function), the introduction ofthat transformed variable to the neural network couldhelp the network learn faster. However, such functionalrelationships are not known in this case.

RESULTS AND IMPLICATIONS FORGENERAL RESERVOIR-QUALITY

PREDICTION

Mineral Distribution and Structural Influences

The mineral distribution in the Zechstein 2 Carbon-ate reflects a general division between dolomites in theplatform deposits, dolomites plus calcites in the slopedeposits, and calcites in the basinal deposits (Figure 2).However, departures from this broad pattern providedclues for development of a geologic model for calcitiza-tion. In particular, the platform deposits contain severalrelatively small areas with distinctively high percent-ages of calcite, closely corresponding to the positions of“bald highs”—tectonically high areas on the platform,bounded by deep-rooted fault systems, where the Zech-stein 2 Carbonate is absent due to removal during Cre-taceous tectonic activity. This relationship provided the

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 51

Page 69: Reservoir Quality Prediction in Sand and Carbonates

52 Love et al.

working hypothesis that calcitization of dolomite wasrelated to fault and fracture systems. The large volumesof the calcium-sulfate–saturated waters required for thecalcitization of dolomite can move from surroundinganhydrite formations through dolomitic units via frac-tures and other permeable zones (Clark, 1980; Warren,1991), providing a mechanism for calcitization of theZechstein 2 Carbonate. Calcitization of dolomite hasbeen linked to fractures and faults in several formations,whether associated with evaporites or not, including theMississippian Madison Limestone (Budai et al., 1984)and the Cretaceous Edwards Limestone (Abbott, 1974).Although this geologic model highlights a plausiblecause-and-effect relationship regarding the distributionof calcite, it is ineffectual in predicting the geographicdistribution of calcite at a pragmatic (fine) scale. Suchpredictions would require detailed paleofault andpaleofracture data that are not available. Consequently,other predictor variables were sought, keeping in mindthat structure was an important factor that might some-how be incorporated into predictive models.

In the Zechstein 2 Carbonate, the same facies havebeen subjected to different degrees of calcitization in dif-ferent parts of the study area, likely reflecting differ-ences in the access of calcitizing fluids to the areas. Forthis reason, formulating one equation to predict calcitefor the entire area produced inadequate results. The for-mation thus was divided into subareas (Figure 3) thatwere delimited based on “structural homogeneity”; thatis, because structure presumably played a substantialrole in influencing the movements of diagenetic fluids,subareas enclose deposits that experienced similar

structural histories (e.g., a spur-and-graben subarea,and a structurally complex, folded and faulted subarea).For each of these structural subareas, a separatemineralogy-predicting equation was generated usingonly data from wells located within that subarea. Table1 indicates the number of wells and samples for each ofthese subareas. The existence of borderlines betweentwo subareas creates the possibility that predictions forone location can be made using two equations (one foreach of the subareas) and, furthermore, that those twopredictions might be meaningfully different. This waschecked by comparing values generated for one locationby the two different equations; in most instances, thevalues matched closely. After initially defining ten sub-areas, predictions were requested for a location at thejuncture of several subareas. To check the predictions(which were extrapolations, because the location wasoutside the well control for each subarea), a new sub-area was defined with the desired prediction locationnear the center (dashed outline in Figure 3, Area I). Thepredictions generated from this model were very simi-lar to those generated by the existing models, althoughthe new model (which does not represent a structurallydefined area) did a poorer job of predicting withhelddata than did the models for surrounding areas.

Porosity and permeability regression equationsand neural network models were generated for eachof the subareas defined during development of themineralogy-predicting equations. The same subareaswere deemed useful because calcitization is the dom-inant control on porosity and permeability; thus,

Figure 3. Map showing sub-division of area into “struc-turally homogeneous”subareas. Black dots repre-sent well locations.

Page 70: Reservoir Quality Prediction in Sand and Carbonates

subareas defined by calcitization controls would besimilar to those defined by porosity and permeabilitycontrols.

Predictor Variables for Calcite, Porosity, and Permeability

Predicting the distribution of reservoir qualityrequires variables whose values are known prior todrilling. For the Zechstein 2 Carbonate, depositionalfacies and location variables meet this requirement;thus, equations to predict calcite, porosity, and perme-ability were developed using only facies and locationvariables and their transformations (e.g., logarithms).These predictor variables were used for both theregression and neural network models. Although notavailable or not as useful for the Zechstein 2 Carbon-ate, formation thicknesses and seismic attributes maybe good predrill predictor variables.

Depositional FaciesFor the Zechstein 2 Carbonate, depositional facies

provide important information about mineralogy,porosity, and permeability due to both depositional anddiagenetic differences among facies. The depositionalfacies characterized in this formation are the platform,platform-lowstand-wedge, upper slope, middle slope,lower slope, and basin. A detailed depositional

framework and sequence stratigraphic model(Strohmenger et al., 1993) had been constructed prior tothis reservoir-quality work, so that the succession offacies could be predicted ahead of drilling a well.

LocationLocation variables (x-y coordinates and depth) play a

crucial role in predicting three-dimensional distribu-tions of parameters. For the Zechstein 2 Carbonate, loca-tion variables significantly predict mineralogy, porosity,and permeability. In addition, the predrill values for xand y are known, and a reasonable estimate can be madefor depth by combining well log and seismic data. Depthunits are in meters, and, for convenient use in the mod-els, the x-y coordinates (e.g., 3452555.0, 5810250.5) weredivided by 106. In addition, as described in the Methodssection, transformations of location variables were usedextensively for the regression models.

As with facies, the mathematical functions of the loca-tion variables that best predict mineralogy, porosity, andpermeability differ in different subareas. When makinga prediction for a facies in one subarea, location providesinformation about smaller-scale areal differences inreservoir-quality distribution; for example, calcite is notdistributed uniformly throughout the upper slopewithin one subarea, but rather may be concentrated inthe upper, eastward portion of that facies.

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 53

Table 1. Porosity and Permeability Prediction Results of Regression and Neural NetworkModels for the Zechstein 2 Carbonate for Withheld Test Data Sets.*

Porosity Prediction PermeabilityAccuracy Prediction Accuracy

Number Numberof Wells of Samples % Within % Within % Within % Within

Subarea† in Subarea in Subarea ±2 porosity % ±4 porosity % ±1 ln(k) md ±1 ln(k) md

A 7 616 38 75 52 8740 60 46 83

B 20 1434 34 62 38 7915 29 38 68

C 10 875 60 76 42 7749 68 35 64

D 23 1807 42 76 55 8534 61 36 73

E 14 908 70 88 64 8857 84 48 73

F 32 5271 42 67 42 7533 56 34 67

G 23 4195 49 75 47 8038 57 44 76

H 32 6990 60 79 49 8045 68 42 73

I 30 3336 42 73 44 7627 53 37 63

J 15 1320 58 87 68 9144 78 52 78

K 81 6925 36 63 41 6826 49 34 60

*Neural network results are in boldface type.†Subareas are outlined in Figure 3.

Page 71: Reservoir Quality Prediction in Sand and Carbonates

54 Love et al.

Calcite Prediction Models

When not used to extrapolate, the regression mod-els used to predict Zechstein 2 Carbonate mineralogyalmost always generate values between 0 and 1, inter-preted as “likelihoods” of calcite (vs. dolomite). Val-ues from 0 to 1 result because the input datum foreach core plug was either a 0 (for dolomite) or a 1 (forcalcite). This binary input was appropriate because, ata core-plug scale, the carbonate samples were almostalways >95% calcite or dolomite. One of the conun-drums of this method lies in interpreting the continu-ous values generated by the discrete models. Because

a dolomite sample is represented by the value 0 and acalcite sample by the value 1, these are the only valuesthat have an unambiguous interpretation. However,when an equation is used to predict calcite ordolomite in an area, a number between 0 and 1 mayresult. If the predicted value for mineralogy is close to0 or 1 (a reasonable definition of close in this casewould be within ~0.3), then the mineralogy reason-ably can be assigned to essentially calcite or essen-tially dolomite. For example, a value of 0.75 can beinterpreted as representing mostly calcite, whereas avalue of 0.25 indicates mostly dolomite. The greaterambiguity arises as to the meaning of an intermediatevalue such as 0.5; this could be interpreted as mean-ing half calcite and half dolomite or as an indetermi-nate mineralogy. Because of this ambiguity, caseswere examined with values near 0.5 to establish a“ground truth” for interpretation. Predicted values of0.5 generally corresponded to wells having both cal-cite and dolomite within the targeted interval. Inpractice, the probability of calcite in a given facies cor-responded closely to the actual percentage of calcitein that facies.

Prediction Equations

Equation 1 is an example of a regression equationused to predict calcite for one structural subareawithin the Zechstein 2 Carbonate. Values for the faciesvariables are 0 or 1, depending on which facies is beingused for the prediction. If, for example, one desires aprediction for the upper slope facies, that variablewould be set to 1, and all other facies variables wouldbe set to 0.

As discussed in the Methods section, all variablesremaining in the predictive equations at the comple-tion of the regression procedure are significant at the0.15 level. The magnitude of the coefficients in theregression equation depends on the magnitude of thevariable associated with that coefficient. Althoughusing location variables and their transforms intro-duces collinearity into regression equations, collinear-ity is not a problem when the equations are used onlyto predict, as they are in this study. However, prob-lems are caused by collinearity if one wants to inter-pret the regression coefficients (e.g., if one wants toknow which variables are most important in eachequation). Again, the equations developed for thisstudy were used only to predict, not to test hypothesesregarding the equations.

Porosity and Permeability Prediction Models

Unlike the mineralogy models, the porosity andpermeability predictive models generate estimates ofporosity and permeability values rather than likeli-hoods. Equation 2 is an example of a regressionequation used to predict porosity. Values for the

Likelihood of Calcite 125.73 3.65( )+

3.51( ) 0.86[ ]0.19(lowstand wedge platform facies)0.23(upper slope facies) 0.38(middle slope facies)

2

= +− +

− ++

– X

Y (ln(depth)

Figure 4. Histograms of residual values (observedvalue minus predicted value) from porosity predic-tion models for one structural subarea (H) withinthe Zechstein 2 Carbonate. The regression modelpredicted 45% of the values within ±2 porosity %,and 68% within ±4 porosity %; the neural networkpredicted 60% of the values within ±2 porosity %,and 79% within ±4 porosity %.

(1)

Page 72: Reservoir Quality Prediction in Sand and Carbonates

facies variables are again 0 or 1, depending on whichfacies is being used for the prediction. As mentioned,the neural network program produces matrices ofweights by which values of the predictor variablesare multiplied to obtain a porosity or permeabilityprediction.

Because depth is present in the equation, the predic-tion is being made for one point (one x,y,z location). Toconstruct a set of horizontal and/or vertical predictions(as along a borehole), the equation is solved numeroustimes, changing x,y,z and facies as appropriate.

Comparison of Neural Network and Regression Results

Two generalities for porosity and permeabilityprediction for the Ca2 emerged during the analyses:porosity is more easily predicted than permeability,and neural networks predict porosity and permeabil-ity better than multiple regression does. The best wayto compare the predictive capabilities of variousmodels is to compare the residuals (i.e., the actualporosity or permeability values minus those pre-dicted by the model) for a set of data not used todevelop the model. Regardless of the goodness-of-fitof any model to the data from which it was gener-ated, the model must be able to predict values fornew data within a reasonable tolerance.

Porosity 2537.7 24.4( ) 403.1( )13.5[ ] 10.9(platform facies)3.9(lowstand wedge platform facies)4.5(upper slope facies) 4.4(middle slope facies)

= + + −+ +

− −−

– X Y2

(ln(depth)

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 55

Figure 5. Plot of predicted vs.observed values for regression andneural network model for one sub-area (H). The vertical groupings ofpoints from the regression analysisrelate to facies, as the contributionsfrom facies must enter the equationadditively. The neural networkplot does not show these group-ings; the predictions are free fromthe linear constraint.

(2)

Page 73: Reservoir Quality Prediction in Sand and Carbonates

56 Love et al.

(A)

(B)

Figure 6. Prediction maps for the top of the Zechstein 2 Carbonate. The maps were made by superimposing a gridover the area with individual cells sized 2.5 × 2.5 km, and generating predictions at every grid x-y intersection.The yellow areas labeled “Bald High” in the legend indicate areas where the Zechstein 2 Carbonate is absent overtectonic features. (A) Mineralogy prediction map. (B) Porosity prediction map.

Page 74: Reservoir Quality Prediction in Sand and Carbonates

Table 1 shows the results for both neural networkand regression model prediction of porosity and per-meability for each of the structural subareas into whichthe Zechstein 2 Carbonate was divided. The columns inTable 1 indicate tolerances (e.g., ± 2 porosity %) about aprediction. The values in Table 1 show the percentagesof withheld porosity and permeability values that werepredicted within a given tolerance. Within each sub-area, the same data were withheld for both the regres-sion and neural network models (the training andcalibration data sets randomly selected by the neuralnetwork model were saved and used for the regressionmodel). The data in Figure 4 show the residual valuesfrom the regression and neural network models for oneof the structural subareas.

Differences exist among the structural subareas withrespect to the ability to predict porosity and permeabil-ity. Subarea B (Figure 3) is one of the most difficultareas in which to make predictions; this may reflect amore complex distribution. In contrast, subarea E mod-els predict quite well, reflecting the simpler porosityand permeability distribution. In general, in subareasfor which porosity prediction is difficult, permeabilityprediction is also a problem.

Table 1 shows that, in general, the neural networkmodels predicted a higher percentage of values withinthe stated tolerance than did the regression models.However, the regression models perform about as wellas the neural network models in predicting average

porosity and permeability values for facies or even inpredicting general trends within a facies (such asdecreasing or increasing porosity). Two pieces of infor-mation are available to help interpret average porosityand permeability values in the case of the Zechstein 2Carbonate: mineralogy predictions and an understand-ing of the geologic variability. If, for example, the min-eralogy predictions indicate a very high probability ofcalcite in the upper slope facies, but the predicted aver-age porosity is high, one would predict that high-porosity dolomite layers exist, because almost all of thecalcite in the upper slope is less than 3% porosity.

If one must predict more detail than that providedby the average, neural network models are recom-mended, as they are when there is a large differencebetween the neural network and regression results orwhere all predictions are poor. Plots of predicted vs.observed values (Figure 5) provide a useful demonstra-tion of the differences in predictive ability between theregression and neural network models for a typicalsubarea. One of the most important differences illus-trated in Figure 5 is that the regression model predictsin “groupings” (visible in Figure 5 as vertical clusters ofpoints), whereas the neural network predictions arecontinuous. The regression clusters relate to the facies;because the regression model is a linear model, theeffects of different facies must enter the model addi-tively—hence the jump in porosity from one facies tothe next. Second, the regression model is constrained in

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 57

Figure 6. (C) Permeability prediction map.

(C)

Page 75: Reservoir Quality Prediction in Sand and Carbonates

58 Love et al.

this case to predicting porosity values no higher than~19%, even though values in the calibration set were ashigh as ~32%. This is again a result of the linear con-straint, which the neural network does not have. Asshown in Figure 5, the neural network model predictedvalues up to ~28%. For some cases, however, the regres-sion models suffice, and have the advantage of relativeease of use, as they can be solved using a calculator.

APPLICATIONS ANDTROUBLESHOOTING

When x-y coordinates and depth are used in predic-tive equations, the equations can generate three-dimensional predictions. Using these equations,three-dimensional representations of the predictionscan be created, and two-dimensional maps or crosssections can be constructed. For the Zechstein 2 Car-bonate, maps were made showing the distribution ofpredicted mineralogy, porosity, and permeability fordifferent horizons. For example, Figure 6 shows mapsof the mineralogy, porosity, and permeability pre-dicted at the top of the formation. Such maps providean overview of the predicted distributions in a givenarea, and constitute a prelude to more detailed predic-tion work, as illustrated in Figure 7 and Table 2.

Because the predictive models for permeabilitywere generated from core-plug permeabilities, theypredict matrix, rather than fracture, permeability.Thus, in areas known to produce from fractures, per-meabilities predicted by the models must be usedwith caution. This limitation can be avoided if reliablewell-log–derived permeabilities are available and if

other desired predictor variables such as facies can beascertained from well logs where no core exists.Another potential problem arises when input values(location, facies) for the equations are not within therange of values used to develop the model (i.e., whenextrapolating rather than interpolating). In such cases,an inconsistent prediction may result, such as a nega-tive porosity. Caution must be exercised when relyingupon extrapolations, even if the answer appears to be“reasonable.”

Although data smoothing led to better predictions,all the models still produced poorly predicted values(e.g., where the observed and predicted values differedby >10 porosity %) (Figure 4). Given this, the questionarises: What, if anything, do the poorly predicted val-ues have in common? For any one model, are thesepoints all from one well or one facies, indicating thatthe model predicts poorly for an entire well location orfacies? Or do the poorly predicted points have nothingin common, simply reflecting a “random” scatteringacross the different wells and facies? If the pointsreflect random scattering, there is a negligible effect onthe ability to predict values representative for a pro-posed well location. If, on the other hand, the poorlypredicted points originate from one area, a separatepredictive model, perhaps using additional predictorvariables, could be tried. This should be checkedbefore applying predictions. After checking many ofthe predictions for the Zechstein 2 Carbonate, the ran-dom scattering scenario was accepted; for example,few wells were found in which the majority of predic-tions within a facies had residual values greater thanabout 4 porosity units.

Figure 7. Example of miner-alogy prediction for an“undrilled” well. The datafrom the well were withheldduring model development,and then a prediction wasmade for the well locationand facies succession. Thecolumn on the left shows thepredicted probability of calcite for each facies; forexample, the upper slopefacies has a 0.75 probabilityof calcite. The column on the right shows the actual mineralogy—the well had acalcite “cap” at the top, andthe remainder was dolomite,as predicted by the lowprobabilities of calciteshown on the left (<0.1).

Page 76: Reservoir Quality Prediction in Sand and Carbonates

Data Requirements for Spatial Predictions

The Zechstein data set used to illustrate the proce-dure for predicting reservoir quality is fairly large, con-sisting of data from 287 wells. However, such a largedata set is not required for making spatial predictions. Infact, because the Zechstein data were divided into sub-areas, some of the predictive models were developedusing fewer than 10 wells (Table 1). Although a generalrule for the necessary density of data cannot be devised,several factors should be considered: (1) number ofwells and number of measurements of the parameter ofinterest, (2) spacing of wells, and (3) complexity of thespatial distribution of the parameter. As the complexityof the distribution increases, more wells at a closer spac-ing are required to achieve useful predictions.

CONCLUSIONS

Optimally, predictive models use reservoir-qualitycontrols, rather than surrogates, as model inputs. How-ever, the three-dimensional distribution of parameterssuch as porosity, permeability, and mineralogy can bepredicted at a pragmatic scale even where cause-and-effect models are not available or do not provide predic-tions at the required scale. For prediction purposes,location variables, along with any other significant vari-ables whose values are known prior to drilling, mayprovide predictive capabilities in statistical models. Prediction of complexly distributed parameters com-monly improves by using neural networks. Overall,predictive models should be judged by their success orfailure, not only by their use of geologic variablesthought to be related to the predicted parametersthrough cause and effect. Such models, however, maynot be reliable for extrapolation purposes.

ACKNOWLEDGMENTS

The authors thank BEB Erdgas und Erdöl GmbH,the W.E.G. publication committee, and Exxon Produc-tion Research Co. for their permission to publish thispaper. Reviews by Tom Jones, Dave Pevear, AltonBrown, and Marek Kacewicz are appreciated. Ananonymous reviewer also contributed comments.

REFERENCES CITED

Abbott, P.L., 1974, Calcitization of Edwards Groupdolomites in the Balcones fault zone aquifer, south-central Texas: Geology, v. 1, p. 359–362.

Bowerman, B.L., and R.T. O’Connell, 1990, Linear sta-tistical models: an applied approach (2d ed.):Boston, PWS-Kent Publishing Co., 1024 p.

Budai, J.M., K.C. Lohmann, and R.M. Owen, 1984, Bur-ial dedolomite in the Mississippian Madison Lime-stone, Wyoming and Utah thrust belt: Journal ofSedimentary Petrology, v. 54, p. 276–288.

Clark, D.N., 1980, The diagenesis of Zechstein carbon-ate sediments; in H. Fuechtbauer and T. Peryt, eds.,The Zechstein Basin with emphasis on carbonatesequences: Contributions to Sedimentology, no. 9, p. 167–203.

Draper, N., and H. Smith, 1981, Applied regressionanalysis (2d ed.): New York, Wiley & Sons Inc., 709 p.

Haykin, S., 1994, Neural networks: New York,Macmillan College Publishing Co., 696 p.

Kacewicz, M., 1994, Model-free estimation of fractureaperture with neural networks: Mathematical Geol-ogy, v. 26, p. 985–994.

Predicting Reservoir Quality Using Linear Regression Models and Neural Networks 59

Table 2. Comparison of Predicted and Actual Porosity for Two Wells.*

Predicted ActualProbability Porosity Porosity

Dep. of Calcite Interpreted Observed (Average) (Average)Well Facies (%) Mineralogy Mineralogy (%) (%)

Plat- Dolomite MostlyLSW 50 and Calcite Calcite 16 15

Upper Mostly Mostly Slope 30 Dolomite Dolomite 11 12

Middle Dolomite DolomiteSlope <10 7 7

Upper Mostly CalciteSlope 70 Calcite 2 2

Middle Mostly MostlySlope 40 Dolomite Dolomite 8 10

Lower Mostly MostlySlope 60 Calcite Calcite 2 2

*Data from these wells were withheld during development of the linear regression equation for this struc-tural subarea. The boldface columns indicate input data (well location and facies). The equation generateda probability of calcite, from which an interpretation of mineralogy was made and compared to theobserved mineralogy observed in the well. In addition, porosity predictions were made.

1

2

Page 77: Reservoir Quality Prediction in Sand and Carbonates

60 Love et al.

Kleinbaum, D.G., and L.L. Kupper, 1978, Appliedregression analysis and other multivariable methods:Boston, Duxbury Press, 556 p.

Strohmenger, C., M. Antonini, G. Jäger, K. Rocken-bauch, and C. Strauss, 1996, Zechstein 2 Carbon-ate reservoir facies distribution in relation toZechstein sequence stratigraphy (Upper Permian,northwest Germany): an integrated approach:Bull. Centre Rech. Explor. Prod. Elf Aquitaine,

v. 20, p. 1–35.Strohmenger, C., K.M. Love, J.C. Mitchell, and K. Rock-

enbauch, 1993, Sedimentology and diagenesis of theZechstein Ca2 Carbonate, Late Permian, NorthwestGermany (abs.): AAPG Bulletin, v. 77, p. 1668.

Warren, J.K., 1991, Sulfate dominated sea-marginaland platform evaporative settings, in J.L. Melvin,ed., Evaporites, petroleum and mineral resources:Developments in Sedimentology 50, p. 69–187.

Page 78: Reservoir Quality Prediction in Sand and Carbonates

Chapter 5

Global Patterns in Sandstone Diagenesis:Their Application to Reservoir QualityPrediction for Petroleum Exploration

ABSTRACT

Sandstones that share common detrital mineralogies, depositional envi-ronments, and burial histories also share common diagenetic histories. Asurvey of the diagenetic history of 100 sandstones from around the worldhas recognized five common, repetitive, and predictable styles of diagenesisin which similar diagenetic mineral assemblages have been observed.

The five diagenetic styles are: (1) quartz, commonly with lesser quantitiesof neoformed clays (e.g., kaolinite and/or illite) and late-diagenetic, ferroancarbonate; (2) clay minerals (illite or kaolinite) with lesser quantities of quartzor zeolite and late-diagenetic carbonate; (3) early diagenetic (low-temperature)grain-coating clay mineral cements such as chlorite, which may inhibit quartzcementation during later burial; (4) early diagenetic carbonate or evaporitecement, often localized, which severely reduces porosity and net pay at veryshallow burial depths; and (5) zeolites, which occur over a wide range in bur-ial temperature, often in association with abundant clay (usually smectite orchlorite) and late-diagenetic, nonferroan carbonates.

61

Primmer, T.J., C.A. Cade, J. Evans, J.G. Gluyas, M.S.Hopkins, N.H. Oxtoby, P.C. Smalley, E.A. Warren,and R.H. Worden, 1997, Global patterns in sand-stone diagenesis: their application to reservoirquality prediction for petroleum exploration, inJ.A. Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 61–77.

Tim J. PrimmerBP Exploration

Dyce, Aberdeen, Scotland, United Kingdom

Chris A. CadeBP Norway Ltd.Forus, Norway

Jonathan EvansBP Exploration

Poole, Dorset, England, United Kingdom

Jon G. GluyasBP Exploration de Venezuela SA

El Rosal, Caracas, Venezuela

Mark S. HopkinsBP Exploration

Sunbury on Thames, England, United Kingdom

Norman. H. OxtobyUniversity of London, Department of Geology

Egham Hill, England, United Kingdom

P. Craig SmalleyEdward A. Warren

BP ExplorationSunbury on Thames, England, United Kingdom

Richard H. WordenDepartment of Geology, Queen’s UniversityBelfast, Northern Ireland, United Kingdom

Page 79: Reservoir Quality Prediction in Sand and Carbonates

62 Primmer et al.

INTRODUCTIONThe quality of a petroleum reservoir is a function of

both its porosity and permeability. In sandstones,these parameters are controlled by initial sedimentcomposition and its subsequent modification duringburial and lithification. Overburden stress acts duringburial to compact the sand, reducing porosity and con-stricting pore throats. Mineral precipitation and disso-lution also affect pore space and may completelychange the pore structure. Extreme, locally intensecementation can also lead to reservoir compartmental-ization, as in the Troll and Murchison fields of theNorth Sea (Gibbons et al., 1993; Prosser et al., 1993).

Accurate assessment of reservoir potential ahead ofdrilling is a critical factor throughout the petroleumexploration and production cycle. In the early stagesof exploration, the limits of economic basement is akey datum to define; the depth at which sedimentsare insufficiently permeable to sustain economicproduction is difficult to define. In a more matureexploration area, accurate prediction of reservoirquality anomalies may become more important,especially where sandstones that are more porousand permeable than would be expected are found,given their burial depth. Prediction of reservoirquality during production tends to concentrate onunderstanding the spatial architecture or localizedheterogeneity in porosity and permeability neededin the course of reservoir management (petroleumproduction and fluid injection).

This chapter presents a pragmatic method for reser-voir quality prediction, one that has been applied suc-cessfully to many oil and gas fields. Given detritalmineral composition, burial depth, and overpressure,the premise of this method is that porosity and perme-ability can be predicted for uncemented sands andcemented sandstones.

A key concept discussed below is “cementationstyle,” the relationship between detrital composition,burial depth, temperature, and cement type. Summa-rized data from published literature and BP’s in-housereports are used to illustrate various styles of diagene-sis, each of which has a specific impact on reservoirquality. Examples of porosity and permeability pre-diction are presented in this volume (Evans et al.;Gluyas; Gluyas and Witton) and elsewhere (Cade etal., 1994; Evans et al., 1994).

PREDICTION OF DIAGENETIC STYLES

Sandstones with similar geological histories (e.g.,sediment composition, depositional environment,facies associations, and burial history) would beexpected to develop similar styles of diagenesis.Although diagenetic cement reduces porosity on a sim-ple volume-for-volume basis, it is important to establisha style containing distinct cements with different habitsand distibutions at the pore scale (e.g., thin pore liningor blocky pore blocking). Similar volumes of differentcements can have dramatically different effects on per-meability. For example, Pallatt et al. (1984) describe thedisproportionate effect that small quantities of authigenic illite can have on permeability. Clearly,establishing a “style” of diagenesis can help provide aframework in which modification of reservoir qualityby postdepositional processes can be predicted morequantitatively. The global database review discussedbelow shows how sediment composition, depositionalenvironment, and burial temperature combine to estab-lish particular styles of diagenesis. The improvedunderstanding based on this analysis can be used topredict likely changes in the porosity and permeabilityof sandstones during burial.

The quartz diagenetic style is the most common and accounts for 40% ofthe sample set. It is also most likely to occur in mineralogically mature sand-stones, while early diagenetic carbonates and zeolites dominate in miner-alogically immature sandstones. Presence or absence of clay appears to beindependent of both initial sand mineralogy and depositional environment.However, when clay is present, the type appears to vary as a function of ini-tial sand mineralogy and depositional environment.

Large quantities of quartz are unusual cements in sequences that havenever been hotter than ~75°C, while illite precipitation at temperaturesbelow ~100°C is rare. Zeolite composition changes systematically fromclinoptilolite at ~25°C to laumonite at temperatures >100°C.

The repetitive nature and simplicity of these five styles can help predictmodifications in reservoir quality due to burial. An accurate prediction ofthe reservoir quality in sandstones forms the basis of an accurate porosityand permeability prediction ahead of drilling wells in petroleum explo-ration, development, or production.

Page 80: Reservoir Quality Prediction in Sand and Carbonates

The Database

The database comprises studies of 100 stratigraphi-cally discrete sandstone units (Figure 1; Table 1). This isnot intended to be an exhaustive review of all availableliterature around the world, but rather a selection ofexamples that are representative of reservoir qualityvariation observed in the regions around the world.Many of the studies reviewed are from North America(mainly the Gulf Coast, the western United States, andCanada). Publications from that part of the world consti-tute a third of the data reviewed. BP’s in-house studies(primarily from the U.K. continental shelf and Porcu-pine Basin) supplement the more-limited open literatureavailable for NW Europe, and form another third of thedata set. The remaining third are studies from otherparts of the world (mainly Africa, South America, andSE Asia). The fact that >80% of the units studied areMesozoic or younger reflects the bias of past work tothose reservoirs that have an economic importance in oiland gas exploration.

The database encompasses a wide range of relevantgeological attributes (depositional environment, sand-stone composition, and maximum burial temperature).The data are dominated by fluvial, deltaic, and shallow-marine sandstones (Figure 2). The relatively small aeo-lian data set probably reflects the poor preservationpotential of this depositional environment. The smallnumber of good deep-marine examples shows howpoorly represented this depositional setting is in someof the well-studied parts of N. America (Wyoming andthe Gulf Coast) and the North Sea. There is also a sig-nificant lack of good descriptions of diagenesis inlacustrine environments. The few examples considered

here, either saline lake or temperate lake deposits, aregrouped with eolian sands or fluviodeltaic deposi-tional environments, respectively.

The compositional maturity of various depositonalenvironments is also shown in Figure 2. Typically,eolian and shallow-marine sands are more maturethan fluvial or deltaic sands, reflecting the degree ofreworking usually encountered in these sorts of depo-sitional environments. In contrast, the compositionalimmaturity of the deep-marine examples may reflectsampling bias, because a significant number of thesestudies are from active volcanic margins (e.g., theWest Coast of the United States), rather than passivemargins or failed rifts sourced by cratonic basement.

Although the selected data are drawn from the mostcomprehensive studies available, data on maximum bur-ial temperature are sparse and often poorly constrained.Estimates of maximum burial temperature were avail-able for just over 60% of the cases studied and rangefrom 25° to 300°C. In an attempt to estimate the effect ofburial temperature (when good field data were not avail-able), fluid inclusion, stable isotope, and organic matura-tion data have been interpreted where appropriate.

RESULTS

Five Styles of Diagenesis in Sandstones

Five common diagenetic styles have been identified(Figure 3). Each has a distinctive diagenetic mineralassemblage. Their characteristics are:

1. Quartz dominated, which often occurs in associa-tion with smaller quantities of neoformed clays

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 63

69

5354

57 6478

4875

39

15

12

4511

31

33

18

40

58

62 63

3235

27

3

6

88

9699

91

98

97

20

74

7,8

10

94

5956

1

4921

79

16

14

52 68

29

89

100

24

62

38

2

4

5

9

13

17,19

22

23

25

28

30

34

36

41 43

42

44

46

47

26

50

51 5565 70

72

60

61 77

6671

7367 76

80

81

82

83 84

85

86

8790

92

93

95

37

Figure 1. Location of reviewed studies.

Page 81: Reservoir Quality Prediction in Sand and Carbonates

64 Primmer et al.

Table 1. List of Studies Reviewed.

Region/Country Formation Age Reference

1 Atlas Mountains, Morocco Adrar N’Dguoe Ordovician Evans, 19902 S. Gabon Rift,

offshore W Africa “Presalt” E. Cretaceous Giroir et al., 19893 Angola Margin,

offshore W Africa “Presalt” E. Cretaceous Girard et al., 19894 W Mali Souroukoto L. Proterozoic Girard and Deynoux, 19915 W Gabon, offshore W Africa (Various) L. Cretaceous Pittman and King, 19866 Niger Delta, Nigeria Agbada Tertiary Lambert and Shaw, 19827 Gulf of Suez, Egypt Rudeis Miocene Evans, 19908 Ras Budran, Egypt Nubian Paleozoic/E. Cretaceous BP in-house9 Ruhuhu Basin, Tanzania Karoo Triassic Wopfner et al., 1990

10 W. Siberia, Russia Vartorsk E. Cretaceous BP in-house11 Alabama Gulf Coast, USA Norphlet L. Jurassic Dixon et al., 198912 N Texas, USA Gray L. Carboniferous Land and Dutton, 197813 Scotian Basin, offshore

E Canada (Various) L. Jurassic/E. Cretaceous Jansa and Urrea, 199014 Wyoming, USA Frontier L. Cretaceous Tillman and Almon, 197915 Louisiana Gulf Coast, USA Woodbine/

Tuscaloosa L. Cretaceous Thomson, 197916 SW Oregon, USA Umpqua Paleocene/Eocene Burns and Etheridge, 197917 N Alaska, USA Sag River L. Triassic/E. Jurassic Mozley and Hoernle, 199018 Wyoming, USA Shannon L. Cretaceous Rangathan and Tye, 198619 N Alaska, USA Nanushuk/Colville E.-L. Cretaceous Smosna, 198820 N Texas, USA Mobeetie L. Carboniferous Dutton and Land, 198521 NE Pacific Coast, USA (Various) Tertiary Galloway, 197922 W Oregon, USA (Various) Eocene Chan, 198523 S Alaska, USA (Various) Jurassic-Paleogene Bolm et al., 198324 California, USA Santa Ynez Paleogene Helmold and Van de Kamp, 198425 Wyoming, USA U. Minnesula E. Permian Market and Al-Shaieb,198426 Mississippi/Alabama, USA Norphlet L. Jurassic McBride et al., 198727 S Ontario, Canada Cataract Silurian O’Shea and Frape, 198828 Texas Gulf Coast, USA Travis Peak E. Cretaceous Dutton and Diggs, 199029 Michigan, USA St. Peter Ordovician Barnes et al., 199130 Wyoming, USA Upper Almond L. Cretaceous Meshri and Walker, 199031 California, USA Stevens Miocene Boles, 198432 N Alaska, USA Kuparak E. Cretaceous Eggert, 198733 Alberta, W Canada Clearwater E. Cretaceous Hutcheon et al., 198934 Texas Gulf Coast, USA Frio Oligocene Milliken et al., 198135 N Alaska, USA Ivishak Permo-Triassic Melvin and Knight, 198436 Alberta, W. Canada Belly River L. Cretaceous Ayalon and Longstaffe, 198837 Wyoming, USA Tensleep Carboniferous Manckiewicz and Steidtmann, 197938 N Alaska, USA Kekiktuk L. Carboniferous Bloch et al., 199039 Texas Gulf Coast, USA Wilcox Eocene Land and Fisher, 198740 Grand Banks,

offshore E Canada Hibernia E. Cretaceous Brown et al., 199041 Wyoming, USA Upper Muddy L. Cretaceous Almon and Davies, 197942 Alberta, W Canada Viking E. Cretaceous Reinson and Foscolos, 198643 Wyoming Lower Muddy L. Cretaceous Almon and Davies, 197944 Offshore Gulf of Mexico (Unspecified) Miocene-Holocene Whynot, 198645 Louisiana Gulf Coast, USA (Unspecified) Plio-Pleistocene Milliken, 198546 N Alaska, USA Kuparak E. Cretaceous BP in-house47 N Mexico Baucarit Miocene Cocheme et al., 198848 C North Sea, UKCS Marnock Triassic Smith et al., 199349 Schleswig-Holstein,

N Germany Dogger M. Jurassic Horn, 196550 Bornholm, Denmark Hasle E. Jurassic Larsen and Friis, 199151 Porcupine Basin,

offshore W Ireland (Unspecified) Permo-Triassic BP in-house52 Barents Shlef, Svalbard Helvetiafjellet E. Cretaceous Edwards, 197953 S North Sea, UKCS Rotliegend E. Permian Gluyas and Leonard, 199554 Irish Sea, UKCS Sherwood E. Triassic Macchi et al., 1990

Page 82: Reservoir Quality Prediction in Sand and Carbonates

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 65

Table 1. (continued.)

Region/Country Formation Age Reference

55 Porcupine Basin, offshore W Ireland (Unspecified) E.-M. Jurassic BP in-house

56 Paris Basin, France Chaunoy Triassic Worden, 199557 N North Sea, UKCS Brae L. Jurassic Gluyas and Coleman, 199258 Haltenbanken, NOCS Garn M. Jurassic Ehrenberg, 199059 Rhine Graben, Germany Bundsandstein Triassic Evans, 199060 Iberian Range, E Spain (Various) Permo-Triassic Morad et al., 199061 Celtic Sea,

offshore SE Ireland Wealden E. Cretaceous BP in-house62 Dorset, S England Bridport E. Jurassic Morris and Shepperd, 198263 Dorser, S England Sherwood E. Triassic Strong and Milodowski, 198764 N North Sea, UKCS Magnus L. Jurassic Emery et al., 199365 Porcupine Basin,

offshore W Ireland (Unspecified) E. Cretaceous Britoil in-house66 S North Sea, offshore

Holland/Germany “J1”–”J4” M.-L. Jurassic BP in-house67 N North Sea, UKCS Brent M. Jurassic Glasmann et al., 198968 Barents Shelf, Svalbard Helvetiafjellet E. Cretaceous Edwards, 197969 E. Midlands, England Crawshaw L. Carboniferous Warren, 198770 Porcupine Basin,

offshore W Ireland (Unspecified) L. Carboniferous BP in-house71 S North Sea, UKCS (Unspecified) L. Carboniferous Cowan, 198972 Porcupine Basin,

offshore W Ireland (Unspecified) M.-L. Jurassic BP in-house73 Inner Moray Firth,

UKCS Beatrice E. Jurassic Haszeldine et al., 198474 E Greenland Vardekloft M. Jurassic BP in-house75 C North Sea, NOCS Ula L. Jurassic Oxtoby et al., 199576 N North Sea, UKCS Statfjord E. Jurassic BP in-house77 Celtic Sea, offshore

SE Ireland Greensand L. Cretaceous BP in-house78 Barents Sea, NOCS Stø E. Jurassic Riches et al., 198679 S Guatemala (Unspecified) Neogene-Holocene Davies et al., 197980 Potiguar Basin,

NE Brazil Pendencia E. Cretaceous Moraes, 199181 N Chile Puilactis L. Cretaceous/Paleocene Hartley et al., 199182 Campos Basin,

offshore SE Brazil Campos L. Cretaceous-Eocene Moraes, 198983 Llanos Basin, Colombia Mirador Eocene Cazier et al., 199584 Llanos Basin, Colombia Guadalup L. Cretaceous Cazier et al., 199585 Huaco, W. Argentina Huaco Neogene Damanti and Jordan, 198986 S Israel Helez E. Cretaceous Shenhav, 197187 Calabria/Sicily, S Italy Stilo-Capo-

d’Orlando Miocene Cavazza and Dahl, 199088 Bengal Basin, Bangladesh Bengal Neogene Imam and Shaw, 198589 Southland Syncline,

New Zealand Murihiku Triassic/Jurassic Boles and Coombs, 197790 Daito Ridge and Basin,

offshore NW Pacific (Unspecified) Eocene Lee, 198891 NW Shelf, offshore Australia Mungaroo M. Jurassic BP in-house92 Gippsland Basin,

offshore SE Australia Latrobe L. Cretaceous/Paleogene Surdam et al., 198993 Gulf of Bohai, N China Shahejie Eocene/Oligocene BP in-house94 E Borneo, Indonesia Mahakan Teriary Rinckenbach, 198895 Yellow Sea, offshore China Fourth and Fifth Paleogene BP in-house96 C Sumatra, Malaysia Sihapas Miocene Gluyas and Oxtoby, 199597 Pattani Basin, Gulf of

Thailand (Unspecified) Miocene Trevena and Clark, 198698 Queensland, E. Australia Surat Cretaceous Hawlader, 199099 S Sumatra, Malaysia Air Benakat Miocene BP in-house

100 N Luzon, Philippines Cagayan Plio-Pleistocene Mathisen, 1984

Page 83: Reservoir Quality Prediction in Sand and Carbonates

66 Primmer et al.

(e.g., kaolinite and/or illite) and late-stage, high-temperature, ferroan carbonate.

2. Clay minerals dominated, such illite or kaolinitewith smaller quantities of quartz or zeolite andlate-diagenetic carbonate.

3. Early diagenetic (low-temperature) grain-coating claymineral cements, such as chlorite. These mayinhibit or restrict subsequent quartz cementationduring burial to higher temperatures. This can,with help from overpressuring, maintain higherporosity than might be expected when buried toconsiderable depths (>3.5 km).

4. Early diagenetic carbonate or evaporite cement domi-nated, often localized, which severely reducesporosity and net pay from very shallow burialdepths.

5. Zeolite dominated, which occur over a wide rangein burial temperatures, often in association withabundant clay (usually smectite or chlorite) andlate-diagenetic nonferroan carbonates.

It is apparent that quartz-dominated diagenesis (rep-resenting 40% of the total) is the most common diage-netic style seen in the selected studies (Figure 4). It is alsonotable that the specific association of early diageneticgrain-coating clay with inhibition of later quartz cementis more common than diagenesis dominated by clayminerals alone. In ~10% of cases, early or late diageneticcarbonates were the predominant cements, and a similarnumber of cases contained significant quantities of zeo-lite. However, evaporite minerals were significantcements in <1% of the examples investigated.

Figure 5 shows the frequency of occurrence of par-ticular clay minerals in the clay-dominated and earlygrain-coating clay with quartz styles of diagenesis.Chlorite is the most commonly occurring clay mineral,mainly as a result of its frequent occurrence as an earlygrain-coating clay. Both kaolinite and illite are lessimportant; they most commonly occur in subordinatequantities with quartz in quartz-dominated diagenesis.Where carbonates occur as significant cements, Fe-cal-cite and Fe-dolomite are the more common late-diage-netic cements, whereas siderite and dolomite are morecommon as early cements (Figure 6). Calcite is com-mon both as an early cement and as a late stage cement.

Effect of Sediment Composition and DepositionalEnvironment on Diagenetic Style

The controls exerted on diagenetic style by differ-ences in primary sediment composition and deposi-tional environment are shown in Figure 7. Quartzcements are more common in sands deposited in envi-ronments where sediment reworking has producedcompositionally more mature (quartzose) sands (e.g.,eolian, deltaic, and shallow-marine environments),whereas sandstones that are mineralogically immatureare likely to be cemented by carbonates and zeolites,regardless of whether sands are arkosic or lithic. On theother hand, there appears to be no apparent correlationbetween sand composition and clay-dominated styles ofdiagenesis. Zeolites seem to be most common in deep-marine sands, but this may reflect the subsample inves-tigated; the majority were from active volcanogenic

0

20

40

60

80

aeolian fluvial fluvio-deltaic

marine-deltaic

shallow marine

deep marine

% arkosic or lithic sandstones

0

5

10

15

20

25

30

No. of studies

Deep marine

Shallow marine

Marine-deltaic

Fluvio-deltaic

Fluvial

Aeolian

Figure 2. Division of selectedstudies by gross depositionalenvironment (top) and com-positional maturity (bottom).All lacustrine examples havebeen grouped in with eithereolian or fluviodeltaic envi-ronments (see text).Compositional maturity ofeach gross depositional envi-ronment is shown in termsof the proportion of arkosicor lithic sands in each depo-sitional environment.

Page 84: Reservoir Quality Prediction in Sand and Carbonates

margins rather than passive margins sourced from cra-tonic basement. Late carbonate cements are more preva-lent (by a factor of 2 or more), but there appears to be nocorrelation between depositional environment and earlycarbonate or clay-dominated styles of diagenesis.

Although no correlation appeared to exist betweenthe occurrence of clay-dominated styles of diagenesisand either sediment composition or depositional envi-ronment, both of these factors influence the type ofdiagenetic clay present in other styles of diagenesis.Kaolinite occurs in more mature (quartzose, sub-arkosic/lithic) sands and is the most common clay inall depositional environments except eolian settings.Illite is more common in quartzose/subarkosic sandsdeposited in eolian or fluvial sands, and chlorite ismost common in deltaic and shallow-marine sands.Chlorite is also the most abundant diagenetic clay inimmature sands, whereas smectite occurs only inimmature sands and is most common in deep-marinedepositional environments.

Sediment composition and depositional environ-ment control the occurrence of some types of earlyand late carbonate cements. Specifically, early sideritecements are most common in relatively mature sandsdeposited in fluvial and marginal-marine settings.This is in contrast to early diagenetic dolomite andearly diagenetic calcite cements, which do not seem to

be correlated with either sediment composition ordepositional environment. Late-diagenetic ferroandolomite cements are most frequently encountered insubarkosic or sublithic sands, whereas late calcitecements are far more abundant in less mature arkosicor lithic sands.

Effect of Maximum Burial Temperatureon Diagenetic Style

Estimates of maximum burial temperature indicatethat quartz cements precipitate over a wide range ofburial temperatures, although fluid inclusion studiessuggest that minimum temperatures of 75°C are usu-ally required for precipitation (Figure 8). Whereasauthigenic clay minerals such as kaolinite and chloriteappear to form over a wide range of temperatures,many studies indicate that illite seems to require sig-nificantly elevated burial temperatures, usually>100°C (Trevena and Clark, 1986; McBride et al., 1987;Cowan, 1989; Girard et al., 1989; Glasmann et al., 1989;Ehrenberg, 1990; Barnes et al., 1992; Emery et al., 1993;Robinson et al., 1993). As noted elsewhere, differentzeolites are stable over relatively narrow temperatureranges in different sedimentary and tectonic environ-ments (Iijima, 1988, and references quoted therein).Temperature estimates for the most commonly

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 67

Quartz

Carbonates

Zeolites

Clays

Evaporites

Common diagenetic minerals

Common styles of

diagenesis

Quartz dominated+ late clays &

carbonate

Clay dominated+ late carbonate &

quartzor zeolite

Early grain coating clays

wholly/partially inhibiting later

quartz+ late carbonate

Early carbonate orevaporite dominated

Zeolite + clays, late carbonate, opal or quartz

Figure 3. Schematic illustration ofthe relationship between common-ly occurring diagenetic cements(shown in order of decreasingabundance) and their associatedstyles.

Page 85: Reservoir Quality Prediction in Sand and Carbonates

68 Primmer et al.

observed zeolites in this review are 15°–85°C forclinoptilolite, 85°–120°C for heulandite, and in >120°Cfor laumonite.

Effect of Cement Import on Diagenetic Style

There is a continuing debate about mass balance(i.e., the extent to which material is supplied orremoved from a sediment during diagenesis) and thedifferent possible sources of cement in particular(Hayes, 1979; Bjørlykke, 1984; Houseknecht, 1988;Gluyas and Coleman, 1992). Answers to the questionof whether sandstones act as open or closed systemsduring burial depend on the size of the system envis-aged. Obviously, on a basin scale, the system is largelyclosed to outside influences, but on the scale of theindividual sandstone pore, the system is open.Between these extremes, at the scale of each strati-graphically distinct sandstone unit, diagenesisappears to be a largely isochemical process, hence thenoted close relationship between sediment composi-tion and diagenetic style. However, in cases where

there is no close relationship between sediment com-position and diagenetic style, some external controlsuch as the import of cementing components from sur-rounding sediments must be invoked. In these cases,the gross depositional environment and sedimentsfrom facies associated with the sandstones underscrutiny become a more significant influence. Forexample, in some clay-dominated styles of diagenesisinvolving illite in compositionally mature eoliansands, import of potassium (among other compo-nents) from associated evaporites appears necessary(McBride et al., 1987; Gluyas and Leonard, 1995).Although reliable data on absolute mineral abun-dances in this review are relatively sparse, it has beenreported elsewhere (Curtis, 1978; Boles, 1981; Gluyas,1985) that cement can be imported to sandstones froma number of sources. These are summarized in Table 2.

Amount of Cement

So far this chapter has considered diagenesis interms of the relative abundance of constituent

0

10

20

30

40

50

quartz dominated

clay dominated

clay with quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

no. of studies

Figure 4. Occurrence of dif-ferent styles of diagenesis intotal data set.

Figure 5. Occurrence of differentclay minerals in clay-dominated orclay with quartz styles of diagenesis.kaolinite

illite

smectite

chlorite

Page 86: Reservoir Quality Prediction in Sand and Carbonates

cements. Although quantitative modal analysis (pointcount) data on mineral abundance exist in the datareviewed, they are of variable quality, vintage, andreliability. This makes consistent comparisons of onestudy with another difficult. However, based on avail-able data, some tentative volumetric ranges can beassigned to each of the styles of diagenesis establishedin Table 3.

To tackle the problem of predicting porosity andpermeability, the diagenetic history of a sandstoneneeds to be reconstructed, and cement volumes need tobe estimated. One approach to this problem is to try tolink the cement abundance range to another variable(in the case of quartz cement, an increase indepth/temperature of burial often corresponds to anincrease in cement volume). Armed with the ranges in

Table 3, a pragmatic approach is to use a “most likely”value within the range tabulated with the ranges them-selves to generate a “most likely” estimate of range inporosity and permeability. This approach is discussedbriefly below and is given in more detail by Gluyas andWitton (this volume).

CONCLUSIONS

The contributions of each of the principal factorscontrolling diagenesis (e.g., sediment composition,depositional environment, burial temperature, andmass import into sandstones) are shown in Figure 9.The present study is not an exhaustive treatment ofclastic diagenesis, but aims to describe the main fac-tors controlling five important styles of diagenesis.

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 69

Figure 6. Occurrence of differentcarbonates in situations whereearly carbonate cement is dominant(top) or where significant late car-bonate occurs (bottom).

Fe-calcite

calcite

dolomite

siderite

Early carbonate

Fe-dolomite

Fe-calcite

calcite

dolomite

siderite

Late carbonate

Page 87: Reservoir Quality Prediction in Sand and Carbonates

70 Primmer et al.

E n t e r P l o t T i t l e

010

20304050607080

quartz

dominated

clay

dominated

clay &

quartz

early

carbonate

dominated

evaporite

dominated

includes

late

carbonate

includes

zeolite

%

010

20304050607080

quartz

dominated

clay

dominated

clay &

quartz

early

carbonate

dominated

evaporite

dominated

includes

late

carbonate

includes

zeolite

%

01020

304050607080

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

01020304050607080

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

0

10203040

5060

7080

quartz

dominated

clay

dominated

clay &

quartz

early

carbonate

dominated

evaporite

dominated

includes

late

carbonate

includes

zeolite

%

DISTRIBUTION OF DIAGENETIC

STYLES FOR SANDSTONES OF

DIFFERENT COMPOSITIONAL

MATURITY

SUBARKOSIC

QUARTZOSE

ARKOSIC

SUBLITHIC

LITHIC

F L

Q

0

10

20

30

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

AEOLIAN

0

10

20

30

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

FLUVIAL

0

10

20

30

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

SHALLOW MARINE

0

102030

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

DEEP MARINE

0

10

20

30

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

MARINE-DELTAIC

0

10

20

30

40

50

60

quartz dominated

clay dominated

clay & quartz

early carbonate dominated

evaporite dominated

includes late

carbonate

includes zeolite

%

FLUVIODELTAIC

0

5

10

15

20

25

30

DISTRIBUTION OF

DIAGENETIC STYLES IN

DIFFERENT GDEs

No.

of s

tudi

es

Figure 7. (A) The influence of different sand compositions on diagenetic style (sediment composition isexpressed in terms of compositional maturity using the scheme of Dott, 1964). (B) The influence of gross depositional environment (GDE) on diagenetic style.

(A)

(B)

Page 88: Reservoir Quality Prediction in Sand and Carbonates

The basic framework of Figure 9 illustrates the different silicate cements that are likely to resultfrom different starting materials at different temper-atures in different depositional environments. Addi-tional parameters are included to show theconditions at which carbonate cements are devel-oped, together with some of the more frequentlyobserved products from material influx into thesandstone.

This chapter has integrated the results of 100studies of diagenesis in sandstones worldwide andestablished a series of regionally consistent patternsof diagenesis. Given a certain minimal amount ofinformation regarding sediment composition, depo-sitional environment, and burial depth and temper-ature, i t seems possible to predict the l ikelydiagenetic history of any sandstone. Although vari-ations in detail from area to area or sandstone tosandstone will exist, and exceptions to the patternsshown in Figure 9 will arise, we expect the findingsoutlined in this review will generally hold true.

THE IMPACT

Prediction of Porosity and Permeability

Besides authigenic cements, the main factor thatinfluences porosity and permeability in sedimentaryrocks is compaction. Compaction curves determinedfrom laboratory experiments enable porosity to be esti-mated as a function of burial depth, overpressure, andductile grain/clay content (Kurkjy, 1988; Gluyas andCade, this volume). These estimates can be furtherrefined by taking into account the most likely diageneticcement predicted at the given depth/temperature ofburial for a particular style of diagenesis in the forma-tion of interest.

Simulations from sphere-pack models (Bryant etal., 1993) have indicated that permeability can be cal-culated directly as a function of porosity, grain size,sorting, and the type of cement present (Cade et al.,1994; Evans et al., 1994). With porosity-depth trendsestablished, analysis of the effect of different styles

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 71

Table 2. External Sources of Cements in Sandstones.

Source At Low Temperatures Supply At Higher Temperatures Supply

Mudrocks Fe2+ for chlorite, SiO2 for quartz,carbon for early carbonates carbon for late carbonates

Carbonates Ca2+ and carbon for early Ca2+ and carbon for late carbonates carbonates

Evaporites Ca2+ for early carbonates K+ for illite,CaSO4 in remobilized evaporites(e.g., anhydrite)

Table 3. Approximate Ranges in Cement Volumes for Different Styles of Diagenesis.

Range in Volume Range in Volume Style of Diagenesis of Principal Cement of Ancillary Cements

Quartz dominated 5–15% 3–5% clay,(increases with <5% late carbonate*temperature of burial)

Clay dominated 10–20% <5% quartz,(only illite dominated <5% late carbonate*increases with temperature of burial)

Early clay/late quartz 5–10% clay, <5% late carbonate*<5% quartz

Early carbonate/ ≤20–30%evaporite dominated (increases in proximity

to evaporites/saline lake deposits)

Zeolite 5–20% ≤10% clay,(increases with ≤10% late carbonate*increasing lithic content)

*Can be locally ≤20–30%, completely occluding porosity.

Page 89: Reservoir Quality Prediction in Sand and Carbonates

72 Primmer et al.

of diagenesis on permeability can be made in termsof a series of characteristic porosity-permeabilitycurves for a given sand grain size and sorting. Twohypothetical examples are shown in Figure 10, inwhich a clean, compositionally mature quartz-cemented sandstone buried to 3000 m is compared toa less mature, subarkosic quartz-kaolinite-illite–cemented sand of similar grain size buried tothe same depth/temperatures of burial. Permeabilityin the subarkosic quartz-kaolinite-illite–cementedsand is consistently lower than the compositionallymore mature quartz-cemented sand at any givenporosity.

Graphical representations such as in Figure 10 canbe used as “maps” during exploration to predict mod-ifications in porosity and permeability relationships asa result of a specific style of the cementation. Predic-tions of overall reservoir effectiveness at any givendepth of burial can be made, allowing a porositythreshold (and, hence, a depth threshold) to be esti-mated from the poroperm relationship established.Specific examples and additional details are givenelsewhere in this volume (Gluyas, this volume; Gluyasand Witton, this volume).

ACKNOWLEDGMENTS

We thank BP Exploration for permission to publishthis paper. We also thank other former members of theReservoir Quality Prediction Team: Kourosh Amiri,

Andrew Brayshaw, Steve Bryant, Dominic Emery,Shona Grant, Andrew Hogg, Clive Maile, Fiona Neall,Joyce Neilson, Hugh Nicholson, and Andrew Robin-son. Thanks also to Mike Bowman, Malcolm McClure,Mark Osborne, Dick Swarbrick, and Sal Bloch for theirconstructive reviews.

REFERENCES CITED

Almon, W.R., and S.H. Davies, 1979, Regional diage-netic trends in the Lower Cretaceous muddy sand-stone, Powder River Basin, in P.A. Scholle and P.R.Schluger, eds., Aspects of diagenesis: SEPM SpecialPublication 26, p. 379–400.

Ayalon, A., and F.J. Longstaffe, 1988, Oxygen isotopestudies of diagenesis and pore water evolution inthe Western Canada sedimentary basin: evidencefrom the Upper Cretaceous Belly River Sandstone,Alberta: Journal of Sedimentary Petrology, v. 58, p. 489–504.

Barnes, D.A., C.E. Lundgren, and M.W. Longman,1992, Sedimentology and diagenesis of the St. PeterSandstone, Central Michigan Basin, United States:AAPG Bulletin, v. 76, p. 1507–1532.

Bjørlykke, K., 1984, Secondary porosity: how impor-tant is it?, in D.A. McDonald and R.C. Surdam, eds.,Clastic diagenesis: AAPG Memoir 37, p. 277–286.

Bloch, S., J.H. McGowen, J.R. Duncan, and D.W. Brizzo-lara, 1990, Porosity prediction, prior to drilling, insandstones of the Kekituk Formation (Mississippian),

Precipitation temperature (°C)

Qua

rtz

cem

enta

tion

dept

h (m

)

-5000

-4000

-3000

-2000

-1000

170150130110907050

HALTENBANKEN

CENTRAL NORTH SEA

MORA Y FIRTH

GULF OF SUEZ

GULF COAST

NORTH SLOPE ALASKA

PARIS BASIN

NORTHERN NORTH SEA

Figure 8. Mean homogeniza-tion temperatures (±1 σ) of fluidinclusions trapped withinquartz cements from sand-stones (from Gluyas et al.,1993). These data are taken toindicate that quartz cementa-tion can occur at any tempera-ture (but over a restrictedtemperature range) above aminimum threshold of 75°C.

Page 90: Reservoir Quality Prediction in Sand and Carbonates

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 73

25°C

Fe-oxidesK-feldspar

PlagioclaseVolcanic glassVolcanic RF

Sedimentary RFMica

K-Feldspar

Microcrystallinequartz

PlagioclaseVolcanic RF

Amorphous silicaAl - smectite

QUARTZFe - SMECTITE

ZEOLITE (CLINOPTILOLITE)ZEOLITE (ANALCITE)

OPAL 75°C

125°

C

Sedimentary RFMica

K-Feldspar

Redistribution ofdetrital quartz

Plagioclase Volcanic RF ZeolitesAl & Fe - smectite

Opal

QUARTZZEOLITE (LAUMONTITE)

ALBITEQUARTZ

CHLORITE

Rea

ctan

tsS

TA

RT

ING

MA

TE

RIA

LS

K-FELDSPAR - RICH ARKOSIC & NON - VOLCANOGENIC

LITHIC SANDS

QUARTZOSESANDS

PLAGIOCLASE-RICH ARKOSIC &

VOLCANOGENIC LITHIC SANDS

Al & Fe - SMECTITEZEOLITE (CLINOPTILOLITE)

AMORPHOUS SILICA

QUARTZ

ILLITE

KAOLINITE

QUARTZ

KAOLINITE

KAOLINITE

External supplyof Fe2+ in fresh

water

CHLORITEOR

SIDERITE

External supplyof highly saline brine

(from evaporites)

ILLITE

K- FELDSPAR

ANHYDRITE NON-FE CALCITE (OR DOLOMITE)

FE - DOLOMITE

External Supply ofCarbonate

CALCITE

CALCITE

Rea

ctan

tsR

eact

ants

Figure 9. Flow chart illustrating the combined control of sediment composition, depositional environment, andburial temperature on diagenetic cements in sandstones.

Page 91: Reservoir Quality Prediction in Sand and Carbonates

74 Primmer et al.

North Slope of Alaska: AAPG Bulletin, v. 74, p. 1371–1385.

Boles, J.R., 1981, Clay diagenesis and the effects onsandstone cementation, in F.J. Longstaffe, ed.: Min-eralogical Association of Canada Short Course inClays and the Resource Geologist, p. 148–168.

Boles, J.R., 1984, Secondary porosity reactions in theStevens Sandstone, San Joaquin Valley, California,in D.A. McDonald and R.C. Surdam, eds., Clasticdiagenesis: AAPG Memoir 37, p. 217–224.

Boles, J.R., and D.S. Coombs, 1977, Zeolite facies alter-ation of sandstones in the Southland Syncline, NewZealand: American Journal of Science, v. 277, p. 982–1012.

Bolm, J.G, T.H. McCulloh, and R.J. Stewart, 1983, Diage-nesis of sandstones in the Lower Cook Inlet, Alaska,and its implications for petroleum plays: Journal ofthe Alaska Geological Society, v. 3, p. 25–31.

Brown, D.M., K.D. McAlpine, and R.W. Yole, 1990,Sedimentology and sandstone diagenesis of theHibernia Formation in Hibernia oil field, GrandBanks of Newfoundland: AAPG Bulletin, v. 73, p. 557–575.

Bryant, S.L., C.A. Cade, and D.W. Mellor, 1993, Per-meability prediction from geologic models: AAPGBulletin, v. 77, p. 1338–1350.

Burns, I.K., and F.G. Etheridge, 1979, Petrology anddiagenetic effects of lithic sandstones: Paleoceneand Eocene Umpqua Formation, Southwest Ore-gon, in P.A. Scholle and P.R. Schluger, eds., Aspects

of diagenesis: SEPM Special Publication 26, p. 307–317.

Cade, C.A., I.J. Evans, and S.L. Bryant, 1994, Analysisof permeability controls—a new approach: ClayMinerals, v. 29, p. 491–501.

Cavazza, W., and J. Dahl, 1990, Notes on the temporalrelationships between sandstone compaction andprecipitation of authigenic minerals: SedimentaryGeology, v. 69, p. 37–44.

Cazier, E.C., A.B. Hayward, G. Espinosa, J. Velandia,J.H. Mugnoit, and W.G. Leel, Jr., 1995, Petroleumgeology of the Cusiana Field, Llanos Basin Foothills,Colombia: AAPG Bulletin, v. 79, p. 1444–1463.

Chan, M., 1985, Correlations of diagenesis with sedi-mentary facies in Eocene sandstones, westernOregon: Journal of Sedimentary Petrology, v. 55,p. 322–333.

Cocheme, J.-J., A. Demant, L. Aguirre, and D. Her-mitte, 1988, Heulandite in the sedimentary filling ofthe “basin and range” (Baucarit Formation) of thenorthern Sierra Madre Occidental, Mexico(abridged English version): Compte Rendu Acade-mie de Sciences Paris, v. 307, Serie II, p. 643–649.

Cowan, G., 1989, Diagenesis of Upper Carboniferoussandstones: southern North Sea Basin, in M.K.G.Whateley and K.T. Pickering, eds., Deltas: sites andtraps for fossil fuels: Geological Society of LondonSpecial Publication 41, p. 57–73.

Curtis, C.D., 1978, Possible links between sandstonediagenesis and depth-related geochemical reactions

Figure 10. Example porosity (φ)-permeability trends constructedusing the principles outlined byCade et al. (1994) for two sands ofsimilar grain size but differentdetrital composition that have expe-rienced similar burial histories.Slightly different sorting character-reflecting differences in compositional maturity are alsoassumed. Note that even with thesame degree of compaction- andcementation-related porosity loss,the predicted permeability of thequartzose sands is 30 times largerthan the subarkosic sand.

Porosity (%)

Per

mea

bili

ty (

mD

)

0.01

0.1

1

10

100

1000

10000

100000

0 5 10 15 20 25 30 35

Compaction followingburial to 3000m

10% quartz

5% kaolinite+illite

quartzose

sand

subarkosic

sand

15% quartz

90mD

3mD

ø = 11%

ø = 26%

Note: Grain size (both sands) =175µm Sorting = mws (quartzose sand)

= ms (sub-arkosic sand)

Page 92: Reservoir Quality Prediction in Sand and Carbonates

in enclosing mudstones: Journal of the GeologicalSociety of London, v. 135, p. 107–118.

Damanti, J.F., and T.E. Jordan, 1989, Cementation andcompaction history of synorogenic foreland basinsedimentary rocks from Huaco, Argentina: AAPGBulletin, v. 73, p. 858–873.

Davies, D.K., W.R. Almon, S.B. Bonis, and B.E. Hunter,1979, Deposition and diagenesis of Tertiary–Holocene volcaniclastics, Guatemala, in P.A. Scholleand P.R. Schluger, eds., Aspects of diagenesis: SEPMSpecial Publication 26, p. 281–306.

Dixon, S.A., D.M. Summers, and R.C. Surdam, 1989,Diagenesis and preservation of porosity in Nor-phlet Formation (Upper Jurassic), southernAlabama: AAPG Bulletin, v. 73, p. 707–728.

Dott, R.H., Jr., 1964, Wacke, graywacke and matrix—what approach to immature sandstone classification?:Journal of Sedimentary Petrology, v. 34, p. 625–632.

Dutton, S.P., and T.N. Diggs, 1990, History of quartzcementation in the Lower Cretaceous Travis PeakFormation, East Texas: Journal of SedimentaryPetrology, v. 60, p. 191–202.

Dutton, S.P., and L.S. Land, 1985, Meteoric burial dia-genesis of Pennsylvanian arkosic sandstones, south-western Anadarko Basin, Texas: AAPG Bulletin, v. 69, p. 22–38.

Edwards, M.B., 1979, Sandstone in Lower CretaceousHelvetiafjellet Formation, Svalbard: bearing onreservoir potential of Barents Shelf: AAPG Bulletin,v. 63, p. 2193–2203.

Eggert, J.T., 1987, Sandstone petrology, diagenesis andreservoir quality, Lower Cretaceous Kuparuk RiverFormation, Kuparuk River field, North Slope Alaska(abs.), in I. Taileur and P. Weimer, eds., Alaska NorthSlope geology, volume 1: Pacific Section of the SEPMand the Alaska Geological Society, p. 108.

Ehrenberg, S.N., 1990, Relationship between diagenesisand reservoir quality in sandstones of the Garn For-mation, Haltenbanken, mid-Norwegian continentalshelf: AAPG Bulletin, v. 74, p. 1538–1558.

Emery, D., P.C. Smalley, and N.H. Oxtoby, 1993, Syn-chronous oil migration and cementation in sand-stone reservoirs demonstrated by quantitativedescription of diagenesis: Philosophical Transactionsof the Royal Society of London, v. A344, p. 115–125.

Evans, A.L., 1990, Miocene sandstone provenancerelations in the Gulf of Suez: insights into synriftunroofing and uplift history: AAPG Bulletin, v. 74,p. 1386–1400.

Evans, I.J., 1989, Geochemical fluxes during shale dia-genesis: an example from the Ordovician of theMoroccan High Atlas: Ph.D. thesis, Reading Uni-versity, 262 p.

Evans, I.J., 1990, Quartz dissolution during shale dia-genesis: implications for quartz cementation insandstones: Chemical Geology, v. 84, p. 239–240.

Evans, I.J., S.L. Bryant, and C.A. Cade, 1993, Modellingthe effect of diagenetic cements on sandstone per-meability, in J. Parnell, A.H. Ruffell, and N.R.Moles, eds.: Geofluids ‘93, p. 212–214.

Evans, J., C. Cade, and S. Bryant, this volume, A geo-logical approach to permeability prediction in

clastic reservoirs, in J.A. Kupecz, J. Gluyas, and S.Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69, p. 91–101.

Galloway, W.E., 1979, Diagenetic control of reservoirquality in arc-derived sandstones: implications forpetroleum exploration, in P.A. Scholle and P.R.Schluger, eds., Aspects of diagenesis: SEPM SpecialPublication 26, p. 251–262.

Gibbons, K., T. Hellen, A. Kjemperud, S.D. Nio, and K.Vebbenstad, 1993, Sequence architecture, faciesdevelopment and carbonate-cemented horizons inthe Troll Field reservoir, offshore Norway, in M. Ash-ton, ed., Advances in reservoir geology: GeologicalSociety of London Special Publication 69, p. 1–31.

Girard, J.P., and M. Deynoux, 1991, Oxygen isotopestudy of diagenetic quartz overgrowths from UpperProterozoic quartzites of western Mali, TaoudeniBasin: implications for quartz cementation: Journalof Sedimentary Petrology, v. 61, p. 406–418.

Girard, J.P., S.M. Savin, and J.L. Aronson, 1989, Diage-nesis of the Lower Cretaceous arkoses of the AngolaMargin: petrologic, K/Ar dating and 18O/16O evidence: Journal of Sedimentary Petrology, v. 59, p. 519–538.

Giroir, G., E. Merino, and D. Nahon, 1989, Diagenesis ofCretaceous sandstone reservoirs of the South GabonRift Basin, West Africa: mineralogy, mass transferand thermal diffusion: Journal of SedimentaryPetrology, v. 47, p. 482–493.

Glasmann, J.R., P.D. Lundegard, R.A. Clark, B.K.Penny, and I.D. Collins, 1989, Geochemical evidencefor the history of diagenesis and fluid migration:Brent Sandstone, Heather field, North Sea: ClayMinerals, v. 24, p. 255–284.

Gluyas, J.G., 1985, Reduction and prediction of sand-stone reservoir potential, Jurassic North Sea: Philo-sophical Transactions of the Royal Society ofLondon, v. A315, p. 187–202.

Gluyas, J.G., this volume, Poroperm prediction forreserves growth exploration: Ula Trend NorwegianNorth Sea, in J.A. Kupecz, J. Gluyas, and S. Bloch,eds., Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 201–212.

Gluyas, J., and C.A. Cade, this volume, Prediction ofporosity in compacted sands, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir69, p. 19–28.

Gluyas, J.G., and M.L. Coleman, 1992, Material fluxand porosity changes during sediment diagenesis:Nature, v. 356, p. 52–53.

Gluyas, J.G., and A.J. Leonard, 1995, Diagenesis: of theRotliegend Sandstone: the answer ain’t blowin’ inthe wind: Marine and Petroleum Geology, v. 12, p. 491–497.

Gluyas, J.G., and N.H. Oxtoby, 1995, Diagenesis ashort (2 million year) story—Miocene sandstones ofCentral Sumatra, Indonesia: Journal of SedimentaryResearch, v. A65, p. 513–521.

Gluyas, J.G., A.G. Robinson, D. Emery, S.M. Grant, andN.H. Oxtoby, 1993, The link between petroleum

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 75

Page 93: Reservoir Quality Prediction in Sand and Carbonates

76 Primmer et al.

emplacement and sandstone cementation, in J.R.Parker, ed., Petroleum Geology of Northwest Europe:Proceedings of the 4th Conference, p. 1395–1402.

Gluyas, J.G., and T. Witton, this volume, Poropermprediction for wildcat exploration prospects:Miocene Epoch, Southern Red Sea, in J.A. Kupecz,J. Gluyas, and S. Bloch, eds., Reservoir quality pre-diction in sandstones and carbonates: AAPGMemoir 69, p. 165–178.

Hartley, A., S. Flint, and P. Turner, 1991, Analcime: acharacteristic authigenic phase of Andean alluvium,northern Chile: Geological Journal, v. 26, p. 189–202.

Haszeldine, R.S., I.N. Samson, and C. Cornford, 1984,Quartz diagenesis and convective fluid movement:Beatrice oilfield, UK North Sea: Clay Minerals, v. 19,p. 391–402.

Hawlader, H.M., 1990, Diagenesis and reservoir poten-tial of volcanogenic sandstones—Cretaceous of theSurat Basin, Australia: Sedimentary Geology, v. 66,p. 181–195.

Hayes, J.B., 1979, Sandstone diagenesis—the hole truth,in P.A. Scholle and P.R. Schluger, eds., Aspects ofdiagenesis: SEPM Special Publication 26, p. 127–139.

Helmold, K.P., and P.C. Van de Kamp, 1984, Diageneticmineralogy and controls on albitization and lau-monite formation in Paleocene arkoses, Santa YnezMountains, California, in D.A. McDonald and R.C.Surdam, eds., Clastic diagenesis: AAPG Memoir 37,p. 239–276.

Horn, V.D., 1965, Diagenese und porositat des Dogger-beta-Hauptsandsteines in den OlfeldernPlon-Ost und Preetz: Erdol Und Kohle-Erdgas-Petrochemie, v. 17, p. 249–258.

Houseknecht, D., 1988, Intergranular pressure-solutionin four quartz sandstones: Journal of SedimentaryPetrology, v. 58, p. 228–246.

Hutcheon, I., H.J. Abercrombie, P. Putnam, R. Gardner,and H.R. Krouse, 1989, Diagenesis and sedimentologyof the Clearwater Lake Formation at Tucker Lake: Bul-letin of Canadian Petroleum Geology, v. 37, p. 83–97.

Iijima, A., 1988, Diagenetic transformations of mineralsas exemplified by zeolites and silica minerals—aJapanese view. Part I: zeolitic diagenesis, in G.V.Chillingarian and K.H. Wolf, eds., Diagenesis II:Developments in Sedimentology, v. 43, p. 147–189.

Imam, M.B., and H.J. Shaw, 1985, The diagenesis of neo-gene clastic sediments from the Bengal Basin,Bangladesh: Journal of Sedimentary Petrology, v. 55,p. 665–671.

Jansa, L.F., and V.H.N. Urrea, 1990, Geology and dia-genetic history of overpressured sandstone reser-voirs, Venture gas field, offshore Nova Scotia,Canada: AAPG Bulletin, v. 74, p. 1640–1658.

Kurkjy, K.A., 1988, Experimental compaction studiesof lithic sands: Master’s thesis, University of Miami,Miami, Florida.

Lambert-Aikhionbare, D.O., and H.F. Shaw, 1982, Sig-nificance of clays in the petroleum geology of theNiger Delta: Clay Minerals, v. 17, p. 91–103.

Land, L.S., and S.P. Dutton, 1978, Cementation of aPennsylvanian deltaic sandstone: isotope data: Jour-nal of Sedimentary Petrology, v. 48, p. 1167–1176.

Land, L.S., and R.S. Fisher, 1987, Wilcox Sandstone

diagenesis, Texas Gulf Coast: a regional isotopiccomparison with the Frio Formation, in J. Marshall,ed., Diagenesis of sedimentary sequences: Geologi-cal Society of London Special Publication 36, p.219–235.

Larsen, O.H., and H. Friis, 1991, Petrography, diagenesisand pore-water evolution of a shallow marine sand-stone Hasle Formation, Lower Jurassic, Bornholm,Denmark: Sedimentary Geology, v. 72, p. 269–284.

Lee, Y.I., 1988, Chemistry and origin of zeolites insandstones at DSDP sites 445 and 446, Daito Ridgeand Basin Province, Northwest Pacific: ChemicalGeology, v. 67, p. 261–273.

Macchi, L., C.D. Curtis, A. Levison, K. Woodward, andC.R. Hughes, 1990, Chemistry, morphology and dis-tribution of illites from Morcambe gas field, Irish Sea,offshore United Kingdom: AAPG Bulletin, v. 74, p. 296–308.

MacDonald, H., P.M. Allan, and J.P.B. Lovell, 1987,Geology of oil accumulation in Block 26/28, Porcu-pine Basin, offshore Ireland, in J. Brooks and K.Glennie, eds.: Petroleum Geology of NW Europe, p. 643–651.

Manckiewicz, D., and J.R. Steidtmann, 1979, Deposi-tional environments and diagenesis of the TensleepSandstone, in P.A. Scholle and P.R. Schluger, eds.,Aspects of diagenesis: SEPM Special Publication 26,p. 319–336.

Markert, J.C., and Z. Al-Shaieb, 1984, Diagenesis andevolution of secondary porosity in Upper Minnelusasandstones, Powder River Basin, Wyoming, in D.A.McDonald and R.C. Surdam, eds., Clastic diagenesis:AAPG Memoir 37, p. 367–389.

Mathisen, M.E., 1984, Diagenesis of Plio-Pleistocenenonmarine sandstones, Cagayan Basin, Philip-pines: early development of secondary porosity involcanic sandstones, in D.A. McDonald and R.C.Surdam, eds., Clastic diagenesis: AAPG Memoir37, p. 177–193.

McBride, E.F., L.S. Land, and L.E. Mack, 1987, Diagen-esis of eolian and fluvial feldspathic sandstones,Norphlet Formation, Upper Jurassic, RankinCounty, Mississippi, and Mobile County, Alabama:AAPG Bulletin, v. 71, p. 1019–1034.

Melvin, J., and A.S. Knight, 1984, Lithofacies, diagenesisand porosity of the Ivishak Formation, Prudhoe BayArea, Alaska, in D.A. McDonald and R.C. Surdam,eds., Clastic diagenesis: AAPG Memoir 37, p. 347–366.

Meshri, I.D., and J.M. Walker, 1990, A study of rock-water interaction and simulation of diagenesis in theUpper Almond Sandstones of the Red Desert andWashakie Basins, Wyoming, in I.D. Meshri and P.J.Ortoleva, eds., Prediction of reservoir quality throughchemical modeling: AAPG Memoir 49, p. 55–83.

Milliken, K.L., 1985, Petrology and burial diagenesis ofPlio-Pleistocene sediments, northern Gulf of Mex-ico: Ph.D. thesis, University of Texas at Austin,Austin, Texas, 112 p.

Milliken, K.L., L.S. Land, and R.G. Loucks, 1981, His-tory of burial diagenesis determined from isotopicgeochemistry, Frio Formation, Brazoria County,

Page 94: Reservoir Quality Prediction in Sand and Carbonates

Texas: AAPG Bulletin, v. 65, p. 1397–1413.Morad, S., I.S. Al-Aasm, K. Ramseyer, R. Marfil, and

A.A. Aldahan, 1990, Diagenesis of carbonate cementsin Permo-Triassic sandstones from the Iberian Range,Spain: evidence from chemical composition and sta-ble isotopes: Sedimentary Geology, v. 67, p. 281–295.

Moraes, M.A.S., 1989, Diagenetic evolution of Creta-ceous–Tertiary turbiditic reservoirs: AAPG Bulletin,v. 73, p. 598–612.

Moraes, M.A.S., 1991, Diagenesis and microscopic het-erogeneity of lacustrine deltaic and turbiditic sand-stone reservoirs, Lower Cretaceous, Potiguar Basin,Brazil: AAPG Bulletin, v. 75, p. 1758–1771.

Morris, K.A., and C.A. Shepperd, 1982, The role of clayminerals in influencing porosity and permeabilitycharacteristics in the Bridport Sands of Wytch Farm,Dorset: Clay Minerals, v. 17, p. 41–54.

Mozley P.S., and K. Hoernle, 1990, Geochemistry ofcarbonate cements in the Sag River and Shublik for-mations, Triassic/Jurassic, North Slope, Alaska:implications for the geochemical evolution of for-mation waters: Sedimentology, v. 37, p. 817–836.

O’Shea, K.J., and S.K. Frape, 1988, Authigenic illite inthe Lower Silurian Cataract Group sandstones ofsouthern Ontario: Bulletin of Canadian PetroleumGeology, v. 36, p. 158–167.

Oxtoby, N.H., A.W. Mitchell, and J.G. Gluyas, 1995, Thefilling and emptying of the Ula oil field, NorwegianNorth Sea, in J.M. Cubitt and W.A. England, eds.,The geochemistry of reservoirs: Geological Society ofLondon Special Publication 86, p. 141–158.

Pallatt, N., M.J. Wilson, and W.J. McHardy, 1984, Therelationship between permeability and the morphol-ogy of diagenetic illite in reservoir rocks: Journal ofPetroleum Technology, v. 36, p. 2225–2227.

Pittman, E.D., and G.E. King, 1986, Petrology and forma-tion damage control, Upper Cretaceous sandstone,offshore Gabon: Clay Minerals, v. 21, p. 781–790.

Prosser, D.J., J.A. Dawes, A.E. Fallick, and B.P.J.Williams, 1993, Geochemistry and diagenesis ofstratabound calcite cement layers within the RannochFormation of the Brent Group, Murchison Field,North Viking Graben (Northern North Sea): Sedi-mentary Geology, v. 87, p. 139–164.

Rangathan,V., and R.S. Tye, 1986, Petrography, diage-nesis and facies control on porosity in ShannonSandstone, Hartzog Draw Field, Wyoming: AAPGBulletin, v. 70, p. 56–69.

Reinson, G.E., and A.E. Foscolos, 1986, Trends in sand-stone diagenesis with depth of burial, Viking For-mation, southern Alberta: Bulletin of CanadianPetroleum Geology, v. 34, p. 126–152.

Riches, P., I. Traub-Sobott, W. Zimmerie, and U.Zinkernagel, 1986, Diagenetic peculiarities of poten-tial Lower Jurassic reservoir sandstones, Troms 1area, off northern Norway, and their tectonic signifi-cance: Clay Minerals, v. 21, p. 565–584.

Rinckenbach, T., 1988, Diagenese minerale des sedi-ments petroliferes du delta fossile de la Mahakam:Ph.D. thesis, L’Universite Louis Pasteur, Strasbourg.

Robinson, A.G., M.L. Coleman, and J.G. Gluyas, 1993,The age of illite cement growth, Village Fields area,southern North Sea: evidence from K-Ar ages and18O/16O ratios: AAPG Bulletin, v. 77, p. 68–80.

Shenhav, H., 1971, Lower Cretaceous sandstone reser-voirs, Israel: petrography, porosity, permeability:AAPG Bulletin, v. 55, p. 2194–2224.

Smith, R.I., N. Hodgson, and M. Fulton, 1993, Salt con-trol on Triassic reservoir distribution, UKCS Cen-tral North Sea, in J.R. Parker, ed., Petroleumgeology of Northwest Europe: Proceedings of the4th Conference, p. 547–558.

Smosna, R., 1988, Low-temperature, low-pressure dia-genesis of Cretaceous sandstones, Alaskan NorthSlope: Journal of Sedimentary Petrology, v. 58, p. 644–655.

Strong, G.E., and A.E. Milodowski, 1987, Aspects ofthe diagenesis of the Sherwood Sandstones of theWessex Basin and their influence on reservoir char-acteristics, in J. Marshall, ed., Diagenesis of sedi-mentary sequences: Geological Society of LondonSpecial Publication 36, p. 325–337.

Surdam, R.C, L.J. Crossey, E.S. Hagen, and H.P.Heasler, 1989, Organic-inorganic interactions andsandstone diagenesis: AAPG Bulletin, v. 73, p. 1–23.

Thomson, A., 1979, Preservation of porosity in the deepWoodbine-Tuscaloosa trend, Louisiana: Gulf CoastAssociation of Geological Society Transactions, v. 30, p. 396–403.

Tillman, R.W., and W.R. Almon, 1979, Diagenesis ofthe Frontier Formation offshore bar sandstones,Spearhead Ranch field, Wyoming, in P.A. Scholleand P.R. Schluger, eds., Aspects of diagenesis:SEPM Special Publication 26, p. 337–378.

Trevena, A.S., and R.A. Clark, 1986, Diagenesis ofsandstone reservoirs of Pattani Basin, Gulf of Thai-land: AAPG Bulletin, v. 70, p. 299–308.

Warren, E.A., 1987, The application of a solution-mineral equilibrium model to the diagenesis of Car-boniferous sandstones, Bothamsall oil field, EastMidlands, England, in J. Marshall, ed., Diagenesisof sedimentary sequences: Geological Society ofLondon Special Publication 36, p. 55–69.

Whynot, J.D., 1986, Mineralogy and early diagenesis ofdeep Gulf of Mexico Basin sediments: Ph.D. thesis,Texas A&M University, 112 p.

Wopfner, H., S. Markwort, and P.M. Semkiwa, 1990,Early diagenetic laumontite in the Lower TriassicManda beds of the Ruhuhu Basin, southern Tanzania:Journal of Sedimentary Petrology, v. 61, p. 65–72.

Worden, R.H., 1996, Carbonate cements in the Triassicsandstones of the Paris Basin, France: origin andeffects.

Global Patterns in Sandstone Diagenesis: Their Application to Reservoir Quality Prediction for Petroleum Exploration 77

Page 95: Reservoir Quality Prediction in Sand and Carbonates

79

Chapter 6

Burial History and Porosity Evolution ofBrazilian Upper Jurassic to Tertiary

Sandstone ReservoirsCristiano Leite Sombra

PETROBRÁS, Centro de Pesquisa e DesenvolvimentoRio de Janeiro, RJ, Brazil

Chang, Hung KiangUniversidade Estadual Paulista, Instituto de Geociências e Ciências Exatas

São Paulo, Sao Paulo, Brazil

ABSTRACT

The parameter time-depth index (TDI) is applied in this study to quantifyempirically the influence of burial history on sandstone porosity evolution.The TDI, expressed in kilometers per million years of age, is defined as thearea in the burial history diagram enclosed by the burial curve of the reser-voir and the axes of the diagram. In practice, reservoir depths during burialhistory are integrated at regular time intervals of 1 m.y. The calculationsexclude present-day bathymetry or paleobathymetry.

Sandstone reservoirs from several sedimentary basins along the Braziliancontinental margin (Santos, Campos, Espírito Santo, Cumuruxatiba,Recôncavo, Sergipe, Alagoas, and Potiguar) were analyzed to investigate theevolution of porosity against TDI. These Upper Jurassic to Tertiary sand-stones lie in depths of 700 to 4900 m, and are hydrocarbon charged (oil orgas). Average porosities of most of these reservoirs were obtained from coreanalysis, and a few porosity data were taken from well log interpretations.Detrital constituents of the sandstones are mainly quartz, feldspar, andgranitic/gneissic rock fragments. Sandstones were grouped into three mainreservoir types, based on composition (detrital quartz content) and grainsorting: Type I (average quartz content <50%) are very coarse grained to con-glomeratic, poorly to very poorly sorted lithic arkoses. Rock fragments aremainly granitic/gneissic and coarse grained. Type II (average quartz contentranging from 50% to 70%) are fine- to coarse-grained (pebbles absent oroccurring in small percentages), moderately sorted arkoses. Type III (averagequartz content >80%) are fine to coarse, moderately to poorly sorted quartzarenites or subarkoses.

Sombra, C.L., and H.K. Chang, 1997, Burial historyand porosity evolution of Brazilian Upper Jurassicto Tertiary sandstone reservoirs, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality pre-diction in sandstones and carbonates: AAPGMemoir 69, p. 79–89.

Page 96: Reservoir Quality Prediction in Sand and Carbonates

80 Sombra and Chang

INTRODUCTION

The initial (depositional) porosity of sandstonesdepends mainly on their grain sorting (Beard and Weil,1973). In the first stages of burial, porosity is mainlyreduced by mechanical compaction. At intermediate toadvanced burial stages, porosity changes are mainlygoverned by chemical reactions (pressure solution,cementation, and dissolution). The bulk effect of thesemechanical and chemical events results in generaltrends of decreasing porosity with increasing depth.Perturbations in such general trends may be introducedby many different parameters, such as framework com-position, early and late cementation, clay coatings, dis-solution, pore fluid composition, pressure (Nagtegaal,1980), geothermal gradient (Galloway, 1974), time-temperature exposure (Schmoker and Gautier, 1988),and duration of burial (Scherer, 1987; Bruhn et al., 1988).

The importance of time during the evolution of reser-voir quality points to a kinetic control on the evolution

of porosity. This has been observed both in laboratoryexperiments and in subsurface data sets. De Boer (1976)concluded, after simulating porosity reduction inquartz-rich sandstone in the laboratory as a function ofpressure, temperature, time and pore fluid (Figure 1),that: (1) porosity decreases with increasing pressure,temperature, and time; (2) if the pore fluid is oil, theporosity reduction is slightly smaller than if the porefluid is water; and (3) time, alone, can account for poros-ity reduction even if temperature and pressure are keptconstant. Siever (1983) suggested that relationshipsamong burial histories, thermal regimes, and rates ofdiagenetic reactions could be compared with petrologicinformation to deduce at what stage in its history a sedi-ment would have accumulated sufficient time and ther-mal energy to accomplish a given extent of reaction.Franks and Forester (1984) proposed that the occurrenceof CO2 in dissolved gases in the Gulf Coast was kineti-cally controlled. Dutta (1986) estimated the kinetic para-meters for the smectite-illite transformation based on

Plots of average porosity against depth show great dispersion in porosityvalues; such dispersion is mostly due to differences in the reservoir burialhistories. However, plotting porosity values against the TDI for individualreservoir types produces well-defined trends. The decrease in porosity is lessmarked in Type III reservoirs, intermediate in Type II, and faster in Type I.Such plots suggest that it is possible to make relatively accurate porosity pre-dictions based on reservoir TDI, texture, and composition, within the con-straints of reservoir depth/age and basin tectonics analyzed in this study.

100

40

36

32

0 40 80 120 160

300 500 500

Pressure (atm)

Time (days)

Po

rosi

ty (

%)

25oC 1 m NaCl Solution

200oC 1 m NaCl Solution

200oC Oil

Figure 1. Experimental sim-ulations of porosity varia-tions as a function of time,temperature, pressure, andpore fluid (de Boer, 1976).

Page 97: Reservoir Quality Prediction in Sand and Carbonates

subsurface data. Other investigators have studied theimportant by-product reactions of the smectite-illitetransformation, such as the generation of organic acids(Crossey et al., 1986) and cementation (Boles, 1978).Schmoker and Gautier (1988) suggested that sand-stone porosity decreases in the subsurface as a powerfunction of thermal maturity. Bruhn et al. (1988), whoanalyzed porosity–depth trends in sediments of the riftphase of Brazilian basins, observed that offshore reser-voirs were more porous than onshore ones, and sug-gested that these differences were related to differencesin the burial histories. Statistical analysis of the influ-ence of 13 distinct parameters on compaction in basinsof average geothermal gradients led to the conclusionthat the first-order parameters are age (time of burial),detrital quartz content, maximum depth of burial, andsorting (Scherer, 1987). Bloch (1991) pointed out that themost important parameters for empirical prediction ofporosity and permeability were grain size/sorting,detrital composition, and temperature history or pres-sure history, or both. Dixon et al. (1989) interpreted thediagenetic evolution of the deep Norphlet Formation ina time-temperature framework.

Many attempts have been made to simulate diage-netic processes based on a time-temperature scenario.Angevine and Turcotte (1983) simulated pressure solu-tion. Leder and Park (1986) simulated quartz cementa-tion. Surdam et al. (1989) constructed a diageneticmodel based mainly on time-temperature–controlledgeneration or destruction of organic acids. Waples andKamata (1993) modeled porosity reduction as a series ofchemical and physical processes, but they did not rec-ommend the use of their model at its current stage ofdevelopment to predict production characteristics of spe-cific reservoirs. Different points of view are also arisingfrom recent research, bringing new interpretations to thecementation and dissolution events. In the diageneticmodel of Smith and Ehrenberg (1989), temperature-controlled equilibria among feldspar, clay, and carbon-ate minerals control dissolution/precipitation of carbon-ate phases; time does not play an important role. Sombraet al. (1990b) did not detect kinetic control on CO2 occur-rence in natural gases of Brazilian sedimentary basins.Gluyas and Coleman (1992) argue that any successfulmodel of cementation by silica must consider source,transportation, and precipitation mechanisms. Numeri-cal models could become possible, but they are hard totest because duration of cementation, and thus flux, isalmost impossible to constrain. Subsequently, Gluyas etal. (1993) examined data from the Garn Formation, Hal-tenbanken, that argue against a direct depth control onquartz cementation, suggesting that cementation tookplace within a restricted time period, associated withrapid subsidence and heating. In this case, origin ofcementation would be associated with a particular time,not a particular temperature or pressure. Either way, itseems that diagenetic models are not well known to thepoint that they can be properly quantified.

The uncertainties in the quantification of diageneticprocesses make empirical models still valuable andoperational tools for porosity prediction. Sombra(1990) defined a new parameter, the time-depth index

(TDI), which reflects time-temperature-pressure expo-sure and can be easily obtained from burial historydiagrams. In this chapter, the TDI is used to estimatethe influence of burial history on sandstone porosityevolution, quantitatively and empirically. The validityof the relationship between porosity and TDI wastested in a data set composed of 38 Late Jurassic to Ter-tiary age sandstone reservoirs of 7 sedimentary basinsalong the Brazilian continental margin (onshore andoffshore). Three main compositional/textural sand-stone types were considered in this study, based ondetrital quartz content and grain sorting.

SANDSTONE RESERVOIRS FROM THEBRAZILIAN CONTINENTAL MARGIN

The main characteristics and the origin of the sand-stone reservoirs from the Brazilian continental marginincluded in this study are described in this section.

General Aspects

Porosity data are mainly from core analysis, whenavailable. In some conglomeratic or unconsolidatedsandstone reservoirs, porosity data are from well loganalysis. Average porosity calculations excluded calciteconcretions, which typically represent <20% of the netpay. In oil/gas fields where the reservoirs have beencored in several wells, a single well was chosen to repre-sent that field. Reservoirs rich in early diagenetic claysintroduced by mechanical infiltration, such as the preriftLate Jurassic fluvial sandstones of the Sergi Formation,were not included. Strongly bioturbated or thin-beddedsandstone/shale sequences were not included in thedata set either. All of the studied reservoirs are hydro-carbon saturated, either oil or gas bearing. Reservoirages range from Late Jurassic to Tertiary. Depths rangefrom 700 to 4900 m (2300–16,000 ft). Temperatures rangefrom 50° to 150°C (122°–302°F), with all basins havingnormal geothermal gradients. Pressures are eitherwithin or close to the normal pressure gradient.

Compositional/Textural Reservoir Types

Sandstone reservoirs were grouped in three maintypes (I, II, and III), based on framework compositionand texture. Sandstones vary from fine grained to con-glomeratic, very poorly sorted to very well sorted;quartz, feldspar, and granitic/gneissic rock fragmentsare the main constituents. Quartz content rangesbetween 40% and 100%. Petrographic analyses used inthis study were taken from previous works (Figure 2):Sombra et al. (1990a) studied Cretaceous marine tur-bidites of the Santos Basin; Moraes (1989) reported onCretaceous and Tertiary marine turbiditic sandstones ofthe Campos Basin; Chang et al. (1983) analyzed Creta-ceous and Tertiary marine turbiditic sandstones of theCumuruxatiba and Espírito Santo basins; Bruhn (1985)studied Cretaceous lacustrine turbiditic sandstones of theCandeias Formation, Recôncavo Basin. Barroso (1987)and Lanzarini and Terra (1989) analyzed Upper Jurassicfluvio-eolian prerift sandstones of the Recôncavo Basin;

Burial History and Porosity Evolution of Brazilian Upper Jurassic to Tertiary Sandstone Reservoirs 81

Page 98: Reservoir Quality Prediction in Sand and Carbonates

82 Sombra and Chang

Abreu (1989) studied Cretaceous lacustrine and marineturbiditic sandstones of the Maceió Formation, Sergipe-Alagoas Basin; Anjos et al. (1990) and Souza (1990)reported on Cretaceous lacustrine fan-deltaic, fluvial,deltaic, and turbiditic sandstones of the Pendencia For-mation, Potiguar Basin; Garcia et al. (1990) studiedUpper Jurassic diagenetic quartz arenites of the SerrariaFormation, Sergipe-Alagoas Basin. Detrital quartz con-tent (measured in percentages), including mono- andpolycrystalline grains, was the parameter determinedto be the main framework composition indicator. Thesethree main reservoir types, called Types I, II, and III, aredescribed as follows:

Type I reservoirs (low quartz content, <50%) arelithic, conglomeratic sandstones that are poorly to verypoorly sorted. Rock fragments are granitic/gneissic.Quartz content of these reservoirs is difficult to ascertainbased on previous petrological works, because samples

for thin sections were biased toward the sandy fractions.Visual estimations of the pebble content from rock frag-ments were made in order to estimate the quartz con-tent. Also included in this reservoir type arefeldspar-rich, medium-grained, moderately sortedsandstones with quartz content <50%. Type I reservoirsinclude mostly apron and fan-deltaic deposits.

Type II reservoirs (intermediate quartz content,50%–70%) are fine- to coarse-grained arkoses that aremoderately to poorly sorted. Pebbles are absent oroccur in low content. Type II reservoirs represent 58%of the data set in this study and include mostly mas-sive slope/basin turbidites in addition to fluvial,deltaic-lacustrine, and fan-deltaic deposits.

Type III reservoirs (quartz content >80%) are fine-to coarse-grained subarkoses or quartz arenites thatare moderately to well sorted. This type includeseolian sands or diagenetic quartz arenites.

TERT.

CRETACEOUS

Q

QQ

Q QQ

QQ

F

F FF

FF

F L

L

LL

L

L

L F L

CAMPOS BASIN(Moraes, 1989)

NORTHEASTERNRECÔNCAVO BASINGOMO Mb.(Bruhn, 1985)

RECÔNCAVO BASINÁGUA GRANDE Mb.(Barroso, 1987)

EOLIAN FACIES OF SERGI Fm.(Lanzarini & Terra, 1989)

SERGIPE-ALAGOAS BASINSERRARIA Fm. IN ROBALO OIL FIELD(Garcia , 1990)

SERGIPE-ALAGOASBASINMUR./MACEIÓ Fm.(Abreu,1989)

SANTOS BASINMERLUZA FIELD(Sombra ,1990a)

CUMURUXATIBA BASINURUCUTUCA Fm.(Chang et al.,1983)

ESP. SANTO BASIN URUCUTUCA Fm.(Chang et al.,1983)

POTIGUAR BASINUPANEMA OIL FIELD(Anjos ,1983)

et al.

et al.

et al.

( Avg. composition)

Type II Reservoirs

Type III Reservoirs

Type I Reservoirs

Figure 2. Detrital composition of sandstone reservoirs of the Brazilian continental margin. F = feldspar, L =lithic fragments, Q = quartz.

Page 99: Reservoir Quality Prediction in Sand and Carbonates

Burial History

The Brazilian offshore sedimentary basins are con-sidered to have originated during regional exten-sional tectonics, with the breakup of Gondwanaresulting in the separation of South America andAfrica (Ponte and Asmus, 1978; Chang et al., 1988).Chang et al. (1988) showed how the stratigraphic evo-lution of Brazilian offshore basins fit into the model ofbasin development of McKenzie (1978). This uniformextension model has two stages of development,which can be summarized as: (1) crustal thinning as aconsequence of stretching of the lithosphere, followedby a passive upwelling of hot asthenosphere, which isresponsible for the initial rift subsidence; and (2) sub-sequent cooling of the lithosphere, which will furtheramplify the initial subsidence, producing thermal orpostrift subsidence. This model explains the burialhistories at the offshore Brazilian continental marginbasins that display two main sedimentation phases:Lower Cretaceous continental sedimentation associ-ated with the rifts, followed by a evaporitic and tran-sitional Aptian deposition, which underlies theCretaceous to Recent open-marine sediments of thepostrift stage. Sedimentation rate in the rift phase wascontrolled mainly by the degree and rate of extensionof the lithosphere. Further subsidence, in addition tothermal subsidence, was influenced by climate, sealevel fluctuations, and sedimentary supply, andresulted in local variations that can be found frombasin to basin, or even within one basin.

On the onshore portions of the marginal basins, ini-tial subsidence associated with the rift stage predomi-nates, with insignificant thermal subsidence. Thesedimentary record is composed almost entirely ofLower Cretaceous continental deposits (lacustrine andfluvial deposits). In those areas, the crustalextension/thinning occurred mainly in the crust, suchas predicted by the nonuniform extension models ofRoyden and Keen (1980) and Wernicke (1985). Becauseof the differences in burial histories, if we compareonshore and offshore reservoirs that lay today at similardepths, the onshore ones were buried first (Figure 3).

Porosity vs. Depth Relationship

There is no clear relationship between averageporosity and depth in the reservoirs that compose thedata set in this study (Figure 4). Even after analyzingspecific reservoir Types I, II, and III, regression analysisof porosity vs. depth reveal very low correlation coeffi-cients (Table 1). The plot of porosity vs. depth for TypeII reservoirs (Figures 5, 6), which represent 58% of thedata set, shows the absence of any clear relationshipbetween these variables. However, reservoirs in off-shore wells have systematically higher porosities thanthe ones in onshore wells (Figure 5), and younger reser-voirs have systematically higher porosities than theolder ones (Figure 6). Bruhn et al. (1988), who studiedrift deposits along the Brazilian continental margin,observed that reservoirs were more porous offshorethan onshore, arguing that differences in burial histo-ries were responsible.

TIME-DEPTH INDEX

The TDI of a reservoir, as defined by Sombra (1990),can be calculated in the burial history diagram of anywell. The index corresponds to the area in the diagramenclosed by the burial curve of the reservoir and theaxes of the diagram (shown in Figure 7). The TDI isexpressed in kilometers times million years of age. Inpractice, reservoir depths during burial history areintegrated at regular time intervals of 1 m.y.

It seems that the first attempt to analyze porosityevolution against an integration of burial depth wasthat from Block et al. (1986). They integrated the burialdepth, for six wells in the Haltenbanken area, withinthe pressure solution domain, i.e., the product of burialtime and depth between 1525 m (5000 ft) and present-day depth. For more details, also see Bloch (1994).

The calculation of the TDI ignores the present-daybathymetry and paleobathymetry. Water depth doesaffect compaction in most situations. The effectivestress along grain-to-grain contacts can be defined asthe vertical stress (total weight of the overburden; i.e.,sediment/fluid plus water column above the reservoir)

Burial History and Porosity Evolution of Brazilian Upper Jurassic to Tertiary Sandstone Reservoirs 83

150 100 50 0 150 100 50 0 150 100 50 0

(a) (b) (c)

ONSHOREBASINS

OFFSHOREBASINS

Age (Ma)

DE

PT

H

Figure 3. Typical schematicburial history diagrams ofonshore (a) and offshore(b and c) reservoirs from theBrazilian continental mar-gin. If we compare onshoreand offshore reservoirs thatlie today at similar depths,onshore reservoirs wereburied first.

Page 100: Reservoir Quality Prediction in Sand and Carbonates

84 Sombra and Chang

minus fluid pressure (Terzaghi and Peck, 1967). Anincrease in effective stress due to water column will becounterbalanced by an equivalent increase of reservoirfluid pressure in reservoirs with approximately normalpressure gradients, or with hydrostatic communication.

The TDI represents a simplification of the procedureproposed by Bloch et al. (1986) and Bloch (1994). Pres-sure solution is probably a kinetically controlledprocess and, in many areas, the pressure-solutiondomain may not be well known. The TDI also repre-sents a simplification of that proposed by Schmokerand Gautier (1988) to predict porosity evolution. Thethermal parameter, which is important because itaffects the mechanical strength of the grains and the

susceptibility for chemical transformations, such aspressure solution, has not been incorporated into thecalibration. The omission stems from the desire to main-tain the relationship as simple and as operational as pos-sible by eliminating parameters that introduceuncertainties (because the reservoirs in the study havebeen deposited in the same tectonic context). The ther-mal parameter is intrinsically present in the TDI para-meter. The TDI reflects the evolution of reservoir depthduring its burial, so it is a number that contains datarelated to the evolution of effective pressure and tem-perature. Vitrinite reflectance is kinetically controlled,and there is good correlation between Brazilian reser-voirs’ TDI and vitrinite reflectance in the associatedshales (Figure 8).

The TDI was calculated for all the reservoirsincluded in this study. The plots of average porosityvs. TDI for the three reservoir types defined (I, II, andIII) show very clear trends of decreasing porosity withincreasing TDI (Figures 9–11). Regression analysis ofporosity on TDI (exponential model) obtained verygood correlation coefficients (Table 2).

Type I reservoirs (lithic, conglomeratic sandstones),which are texturally and compositionally the mostimmature and presented lower initial porosities, alsodisplay a very rapid porosity decay with increasing

0 10 20 300

1

2

3

4

5

TYPE I RESERVOIRSTYPE II RESERVOIRSTYPE III RESERVOIRS

R

R

PR

R SE

P

P

R SE C

CSE

ALAL

R SER

R

C

P

AL

RES

RSE

CPES

P

SE

AL

RR

ES

ESSE

S

S

x

x

xx x

x

x

x

x

x

Porosity (%)D

epth

(km

)

Figure 4. Plot of average porosity of reservoirs stud-ied against depth. AL = Alagoas Basin; C = CamposBasin; ES = Espírito Santo/Cumuruxatiba basins; P =Potiguar Basin; R = Recôncavo Basin; S = SantosBasin; SE = Sergipe Basin.

Figure 5. Plot of average porosity of Type II reser-voirs against depth, for onshore and offshore wells.Reservoirs located offshore are more porous. C =Campos Basin; P = Potiguar Basin; R = RecôncavoBasin; S = Santos Basin; SE = Sergipe Basin.

0 10 20 300

1

2

3

4

5

OFFSHORE WELLSONSHORE WELLS

R

R

P

SE

PP

R SE C

CSE

SE

C

P

P

SEC

P

P

SE

S

S

Porosity (%)

Dep

th(k

m)

TYPE II RESERVOIRS

Table 1. Simple Linear Regression of Porosity onDepth for Reservoir Types I, II, and III.*

Reservoir Type b a n r2(%)

I 11.3 –4.33E–4 8 1.0II 20.2 –8.30E–4 21 1.3III 28.1 –3.88E–3 8 73.3

*Porosity (%) = b + a; Depth (km); n = number of points in data set.

Page 101: Reservoir Quality Prediction in Sand and Carbonates

TDI (Figure 9). The data set for this reservoir type isvery limited, and only cautious conclusions can bemade for TDI values >200 km ×Ma based on observedtrends. However, we expect to find very low porosities(<10%) for TDI values >200 km ×Ma.

Type II reservoirs (feldspar rich) represent the bestdocumented reservoir type (58% of the data set). Agood trend of porosity decline with increasing TDIcan be seen for this reservoir type (Figure 10). A shiftto increased porosity with depth is observed with thecomparison of Type I and Type II reservoirs. Type IIis invariably more porous. Two outlier points weredocumented as a typical case of porosity preservationat great depth due to early chlorite coatings. Sombraet al. (1990a) concluded that the porosity preserveddue to early chlorite coatings was 9% and 4% in thesetwo wells, after comparing chlorite-coated and chlorite-free sandstones.

Type III (quartz-rich) are the most porous reser-voirs, displaying a good trend of porosity decline vs.TDI (Figure 11). However, the data for this reservoirtype are limited, and this trend must be viewed cau-tiously. The chemical and mechanical stability ofquartz is probably responsible for the highest porositypreservation in this reservoir type. Quartz enrichmentin these reservoirs was related to either depositional or

diagenetic processes; quartz-rich sandstones weredeposited in eolian settings. Diagenetic quartz arenitesresulted from the leaching of feldspars and rock frag-ments close to regional unconformities, as in the Ser-raria Formation, Sergipe Basin (Garcia et al., 1990).Porosity destruction due to intense silica cementationtends to be an important diagenetic event in quartz-rich sandstones at elevated temperatures such as100°–150°C (212°–302°F) (Bjørlykke et al., 1989). In suchreservoirs, the presence of hydrocarbons is essential forporosity preservation because they retard diagenesis.In a well from Sergipe Basin, at 4300 m (14,100 ft) theaverage porosity of the Serraria Formation is 15%above the oil-water contact and near zero below thiscontact (Garcia et al., 1990), with quartz cementationresponsible for porosity destruction.

DISCUSSION

Diagenetic processes related to porosity destructionare grouped into either compaction or cementation.Compaction consists of two modes, mechanical andchemical. In the former, porosity reduction is caused bygrain rearrangement in response to applied stress, usu-ally associated with incremental overburden. Chemicalcompaction results from rearrangement of frameworkgrains that underwent chemical dissolution, particu-larly along the regions of major contacts or stress con-centrations. This remobilization produces an additionalreduction of volume compared with pure mechanicalcompaction. Diagenetic and mass balance studies per-formed in siliciclastic sequences of three Brazilianbasins of the equatorial margin (Chang, 1983) led to theconclusion that compaction (mechanical and chemical)is the main diagenetic factor controlling porosity decayduring burial. Lundegard (1992) analyzed a large data-base of point count from diverse sandstones and con-cluded that compaction (mechanical and chemical),although being generally underappreciated, is probablythe dominant mechanism of porosity loss in sandstones.

Burial History and Porosity Evolution of Brazilian Upper Jurassic to Tertiary Sandstone Reservoirs 85

Figure 6. Plot of average porosity of Type II reser-voirs against depth, by age. Younger reservoirs aremore porous. C = Campos Basin; P = Potiguar Basin;R = Recôncavo Basin; S = Santos Basin; SE = SergipeBasin.

Figure 7. Burial history diagram of a hypotheticalreservoir, illustrating the significance of the time-depth index (i.e., area enclosed by the axes and thereservoir burial curve).

0 10 20 300

1

2

3

4

5

LATE JURASSIC/EARLY CRETACEOUSLATE/MID-CRETACEOUSTERTIARY

R

R

P

SE

PP

R SE C

CSE

SE

C

P

R

SEC

P

P

SE

S

S

Porosity (%)D

epth

(km

)

TYPE II RESERVOIRS

X

X

X

XX

X

X

X

Dep

th

Age (Ma)

(km

)

150 100 50 0

2

0

4

6

RESERVOIR BURIALHISTORY

AREA=TIME-DEPTHINDEX

Page 102: Reservoir Quality Prediction in Sand and Carbonates

86 Sombra and Chang

The role of cementation is that of porosity reductionby filling the void space with authigenic mineral pre-cipitation. As reasoned by Bjørlykke et al. (1989), thismode of porosity decrease occurs without loss in thebulk volume, in contrast to the bulk volume reductionthat results from mechanical compaction. Therefore, aporosity trend produced solely by cementation or dis-solution should not significantly affect the commonlyobserved trend of porosity reduction with depth. Thisstatement will hold more strongly if diagenetic trans-formations occurring in the sandstones are relativelyisochemical, as it has been suggested by severalauthors (Chang, 1983; Bjørlykke et al., 1988; Giles andde Boer, 1990).

Compaction is essentially bulk volume reductionthat involves the removal of a fluid phase from a

porous solid. The physics involved in the compactionof two-phase porous media (solid framework andfluid) has been the subject of many reports (Sharpand Domenico, 1976; Bethke, 1985; Nakayama andLerche, 1987; Mello, 1994). Mello (1994) presents anexcellent overview of the rheology of compactingporous sediments on a geological time scale. Sedi-ment rheology must be accounted for to properlymodel porosity reduction. At room temperature,most consolidated sedimentary rocks are brittle,which means that they behave elastically until theyfail. At higher temperature and pressure, buried sed-iments behave like ductile material. Both brittle andductile deformations are permanent and irreversible.Observations of sediments at geological time scaleindicate that sediments exhibit three rheological com-ponents: elastic, plastic, and viscous. The predomi-nance of each behavior is dependent on composition,temperature, state of stress, degree of lithification,and length of time. For instance, sediment deforma-tion on a short time scale responds elastically,because there is not enough time for the fluids to beremoved. A complete rheological model that is geo-logically realistic is an elasto-viscoplastic model(Mello, 1994).

The porosity trend exhibited for the three types ofreservoirs is at least qualitatively consistent with the

**

** ****

** **

** **

****

**

**

**

**

**

**

**

****

0.2 0.4 0.6 0.8 1.0 1.2

200

100

0

300

400

Tim

e-D

epth

Ind

ex(k

Ma)

Vitrinite Reflectance ( Ro% )

Figure 8. Plot of time-depth index of some reservoirsof the Brazilian continental margin against vitrinitereflectance in the associated shales.

Figure 9. Plot of porosity (average and maximum)against time-depth index for Type I reservoirs.Broken lines enclose average porosity values. AL =Alagoas Basin; ES = Espírito Santo/Cumuruxatibabasins; F = feldspar; L = lithic fragments; Q = quartz.

TYPE IRESERVOIRS

QQ

FF LL

AL

ESAL

ES

ALES

ES

AL

Core analysisLog analysis

500

400

300

200

100

0

0 10 20 30

Porosity ( % )

Tim

e-D

epth

Ind

ex(k

Ma)

Table 2. Exponential Regressions of Porosity (%) onTime-Depth Index (km ×Ma) for Reservoir Types I,II, and III.*

Reservoir Type b a n r2(%)

I 3.07 –4.79E–3 8 84.7II 3.34 –3.08E–3 21 82.2III 3.46 –1.92E–3 8 82.9

*Porosity (%) = exp (b + a TDI); n = number of points in data set.

Page 103: Reservoir Quality Prediction in Sand and Carbonates

variables listed above and influential on the rheologi-cal behavior. For instance, Type II or Type III reser-voirs have a higher quartz content, and subsequentlya low content of lithic (ductile) rock fragments. As aresult, the compaction trend shows less porosityreduction. This compositional control dominatesbecause the thermal regimes and burial histories arevery similar.

CONCLUSIONS

Burial history plays a very important role in the evo-lution of sandstone porosity along the Brazilian conti-nental margin, in addition to detrital composition andtexture. Reservoirs that have resided at maximum buriallonger tend to be less porous than the ones that achievedmaximum burial late in their burial history, indicatingtime is a factor in porosity destruction. The more miner-alogically and texturally mature sandstones lose poros-ity at a slower rate than the immature ones duringprogressive burial. The decay in porosity is poorlyrelated to the present-day depth of the sandstone reser-voirs, but it is closely related to the evolution of depthduring burial. Good relationships were obtainedbetween porosity and TDI, a parameter that reflects theevolution of reservoir depth over geologic time.

Porosity prediction of a sandstone reservoir along theBrazilian continental margin is possible with informa-tion about its mineralogy, texture, and burial history.Exceptions observed were sandstones that containedearly chlorite coatings, which preserved porosity. Theranges of detrital composition, texture, age, depth, tem-perature, and burial history of these sandstones shouldbe considered when making porosity predictions.

ACKNOWLEDGMENTS

We thank AAPG reviewers M. Emery and J.Schmoker. We thank J. Gluyas, S. Bloch, and Carlos H.L. Bruhn for suggestions and discussions. We alsothank Sylvia Anjos, Luis F. De Ros, and Rogerio Schifferde Souza for exchanging ideas. We thank PETROBRÁSfor granting permission to publish this paper.

REFERENCES CITED

Abreu, C.J., 1989, Predicting reservoir quality in theCretaceous Maceió Member of the Sergipe-AlagoasBasin, northeast Brazil: Master’s thesis, Universityof Cincinnati, Cincinnati, Ohio, 106 p.

Angevine, C.L., and D.L. Turcotte, 1983, Porosityreduction by pressure solution: a theoretical model

Burial History and Porosity Evolution of Brazilian Upper Jurassic to Tertiary Sandstone Reservoirs 87

Figure 10. Plot of porosity (average and maximum)against time-depth index for Type II reservoirs.Broken lines enclose average porosity values.Outliers had porosity preserved by early chloritecoatings (Sombra et al., 1990a). C = Campos Basin;P = Potiguar Basin; F = feldspar; L = lithic frag-ments; Q = quartz; R = Recôncavo Basin; S = SantosBasin; SE = Sergipe Basin.

Porosity ( % )T

ime

-Dep

thIn

dex

(km

×M

a)

0 10 20 30

100

200

300

400

500

0

TYPE I IRESERVOIRS

Core analysisLog analysis

C

SEC

CP

C

SE SE

SE

SESE

R

PP

R

PS

SP

R

R

Q

LF

P

Porosity (% )

Tim

e-D

epth

Ind

ex(k

Ma

)

0 10 20 300

100

200

300

400

500

TYPE I I IRESERVOIRS

R

R

R

R

R

R

SE

Q

F L

Figure 11. Plot of porosity (average and maximum)against time-depth index for Type III reservoirs.Broken lines enclose average porosity values. F =feldspar; L = lithic fragments; Q = quartz; R =Recôncavo Basin; SE = Sergipe Basin.

Page 104: Reservoir Quality Prediction in Sand and Carbonates

88 Sombra and Chang

for quartz arenites: Geological Society of AmericaBulletin, v. 94, p. 1129–1134.

Anjos, S.M.C., C.L. Sombra, R.S. Souza, and R.N. Waick,1990, Potencial de reservatórios profundos na For-mação Pendência, Bacia Potiguar Emersa: Boletim deGeociências da Petrobrás, v. 4, p. 509–530.

Barroso, A.S., 1987, Diagênese e eficiência de recuper-ação dos reservatórios do Campo de Araças, Baciado Recôncavo, Brasil: Master’s thesis, UniversidadeFederal de Ouro Preto, Ouro Preto, Brasil, 160 p.

Beard, D.C., and P.K. Weil, 1973, Influence of textureon porosity and permeability of unconsolidatedsands: AAPG Bulletin, v. 57, p. 349–369.

Bethke, C.M., 1985, A numerical model of compaction-driven groundwater flow and heat transfer and itsapplication to the paleohydrology of intracratonicsedimentary basins: Journal of GeophysicalResearch, v. 90, p. 6817–6828.

Bjørlykke, K., A. Mo, and E. Palm, 1988, Modelling ofthermal convection in sedimentary basins and itsrelevance to diagenetic reactions: Marine andPetroleum Geology, v. 5, p. 338–351.

Bjørlykke, K., M. Ram, and G.C. Saigal, 1989, Sandstonediagenesis and porosity modification during basinevolution: Geologisches Rundschau, v. 78, p. 243–268.

Bloch, S., 1991, Empirical prediction of porosity andpermeability in sandstones: AAPG Bulletin, v. 75, p. 1145–1160.

Bloch, S., 1994, Case histories––offshore Mid-Norway/Taranaki Basin, New Zeland/San Emigdio area,California, in M. D. Wilson, ed., Reservoir qualityassessment and prediction in clastic rocks: SEPMShort Course 30, p. 357–365.

Bloch, S., R.K. Sucheki, J.R. Duncan, and K. Bjørlykke,1986, Porosity prediction in quartz-rich sandstones:Middle Jurassic, Haltanbanken area, offshore centralNorway (abs.): AAPG Bulletin, v. 70, p. 567.

Boles, J. R., 1978, Active ankerite cementation in thesubsurface Eocene of Southwest Texas: Contribu-tions to Mineralogical Petrology, v. 68, p. 13–22.

Bruhn, C.H.L., 1985, Sedimentação e evolução dia-genética dos turbiditos eocretácicos do MembroGomo, Formação Candeias, no compartimentonordeste da Bacia do Recôncavo, Bahia: Master’s the-sis, Universidade Federal de Ouro Preto, Ouro Preto,Brasil, 203 p.

Bruhn, C.H.L., C. Cainelli, and R.M.D. Matos, 1988,Habitat do petróleo e fronteiras exploratórias nosrifts Brasileiros: Boletim de Geociências da Petrobrás,v. 2, p. 217–254.

Chang, H.K., 1983, Diagenesis and mass transfer inCretaceous sandstone-shale sequences, offshoreBrazil: Ph.D. thesis, Northwestern University,Evanston, Illinois, 339 p.

Chang, H.K., S.M.C. Anjos, and C.R. Drug, 1983, Car-acterísticas dos reservatórios e evolução diagenéticada sequência turbidítica do Cretáceo Superior e Ter-ciário Inferior das Bacias do Espírito Santo e Cumu-ruxatiba: Rio de Janeiro, Petrobrás internal report.

Chang, H.K., R.O. Kowsmann, and M.F. Figueiredo,1988, New concepts on the development of east

Brazilian marginal basins: Episodes, v. 11, p. 194–202.Crossey, L.J., R.C. Surdam, and R.W. Lahann, 1986,

Application of organic/inorganic diagenesis toporosity prediction, in D. L. Gautier, ed., Roles oforganic matter in sediment diagenesis: SEPM Spe-cial Publication 38, p. 147–156.

de Boer, R.B., 1976, Thermodynamical and experimen-tal aspects of pressure solution, in J. Cadek and T.Paces, eds., Proceedings of the International Sym-posium on Water-Rock Interactions: GeologicalSurvey, Prague, 1974, p. 381–387.

Dixon, S.A., D.M. Summers, and R.C. Surdam, 1989,Diagenesis and preservation of porosity in Nor-phlet Formation (Upper Jurassic), southernAlabama: AAPG Bulletin, v. 73, p. 707–728.

Dutta, N.C., 1986, Shale compaction, burial diagenesisand geopressures: a dynamical model, solution andsome results, in J. Burrus, ed., Thermal modeling insedimentary basins: Paris, Editions Technip, Collec-tion Colloques et Seminaires 44, p. 149–172.

Franks, S., and R. Forester, 1984, Relationships amongsecondary porosity, pore fluid chemistry and car-bon dioxide, Texas Gulf Coast, in W.S. McDonaldand R.C. Surdam, eds., Clastic diagenesis: AAPGMemoir 37, p. 63–80.

Galloway, W.E., 1974, Deposition and diagenetic alter-ation of sandstone in a Northeast Pacific arc-relatedbasin: implications for graywacke genesis: Geologi-cal Society of America Bulletin, v. 85, p. 379–390.

Garcia, A.J.V., L.F. de Ros, R.S. Souza, and C.H.L.Bruhn, 1990, Potencial de reservatórios profundos naFormação Serraria, Bacia de Sergipe-Alagoas: Bole-tim de Geociências da Petrobrás, v. 4, p. 467–488.

Giles, M.R., and R.B. de Boer, 1990, Origin and signifi-cance of redistributional secondary porosity:Marine and Petroleum Geology, v. 7, p. 378–397.

Gluyas, J., and M. Coleman, 1992, Material flux andporosity changes during sediment diagenesis:Nature, v. 356, p. 52–54.

Gluyas, J.G., S.M. Grant, and A.G. Robinson, 1993,Geochemical evidence for a temporal control onsandstone cementation, in A.D. Horbury and A.G.Robinson, eds., Diagenesis and basin development:AAPG Studies in Geology 36, p. 23–33.

Lanzarini, W.L., and G.J.S. Terra, 1989, Fácies sedi-mentares, evolução da porosidade e qualidade dereservatório da Formação Sergi, Campo de FazendaBoa Esperança, Bacia do Recôncavo: Boletim deGeociências da Petrobrás, v. 3, p. 365–375.

Leder, F., and W.C. Park, 1986, Porosity reduction insandstone by quartz overgrowth: AAPG Bulletin,v. 70, p. 1713–1728.

Lundegard, P.D., 1992, Sandstone porosity loss—a“big picture” view of the importance of compaction:Journal of Sedimentary Petrology, v. 62, p. 250–260.

McKenzie, D., 1978, Some remarks on the develop-ment of sedimentary basins: Earth and PlanetaryScience Letters, v. 40, p. 25–32.

Mello, U.T., 1994, Thermal and mechanical history ofsediments in extensional basins: Ph.D. thesis,Columbia University, New York, 395 p.

Moraes, M.S., 1989, Diagenetic evolution of Cretaceous–

Page 105: Reservoir Quality Prediction in Sand and Carbonates

Tertiary turbidite reservoirs, Campos Basin, Brazil:AAPG Bulletin, v. 73, p. 598–612.

Nagtegaal, P.J.C., 1980, Diagenetic models for predictingclastic reservoir quality: Barcelona, Revista del Insti-tuto de Investigaciones Geologicas, v. 34, p. 5–19.

Nakayama, K., and I. Lerche, 1987, Two-dimensionalbasin analysis, in B. Doliges, ed., Migration ofhydrocarbons in sedimentary basins: Paris, EditionsTechnip, p. 597–611.

Ponte, F.C., and H.E. Asmus, 1978, Geological frame-work of the Brazilian continental margin: Geologis-ches Rundschau, v. 67, p. 201–235.

Royden, L., and C.E. Keen, 1980, Rifting processes andthermal evolution of the continental margin of east-ern Canada determined from subsidence curves:Earth and Planetary Science Letters, v. 51, p. 343–361.

Scherer, M., 1987, Parameters influencing porosity insandstones: a model for sandstone porosity predic-tion: AAPG Bulletin, v. 71, p. 485–491.

Schmoker, J., and D.L. Gautier, 1988, Sandstone poros-ity as a function of thermal maturity: Geology, v. 16,p. 1007–1010.

Sharp, J.M., Jr., and P.A. Domenico, 1976, Energytransport in thick sequences of compacting sedi-ments: Geological Society of America Bulletin, v. 87,p. 390–400.

Siever, R., 1983, Burial history and diagenetic reactionkinetics: AAPG Bulletin, v. 67, p. 684–691.

Smith, J.T., and S.N. Ehrenberg, 1989, Correlation ofcarbon dioxide abundance with temperature inclastic hydrocarbon reservoirs—relationship toinorganic chemical equilibrium: Marine andPetroleum Geology, v. 6, p. 129–135.

Sombra, C.S., 1990, O papel da história de soterra-

mento na evolução da porosidade de arenitos(bacias marginais Brasileiras): Boletim de Geociên-cias da Petrobrás, v. 4, p. 413–428.

Sombra, C.L., L.M. Arienti, M.J. Pereira, and J.M.Macedo, 1990a, Parâmetros controladores daporosidade e da permeabilidade nos reservatóriosclásticos profundos do Campo de Merluza, Bacia deSantos, Brasil: Boletim de Geociências da Petrobrás,v. 4, p. 451–466.

Sombra, C.L., T. Takaki, G.I. Henz, and A.S. Barroso,1990b, CO2 in natural gases of Brazilian sedimen-tary basins (abs.): AAPG Bulletin, v. 4, p. 768.

Souza, R.S., 1990, Controle deposicional e diagenéticodos reservatórios profundos do Campo de Pescada,Bacia Potiguar: Boletim de Geociências da Petrobrás,v. 4, p. 531–553.

Surdam, R.C., T.L. Dunn, H.P. Heasler, and D.B. Mac-Gowan, 1989, Porosity evolution in sandstone/shalesystems, in I.E. Hutcheon, ed., Burial diagenesis:Mineralogical Association of Canada, Short CourseHandbook, v. 15, p. 61–134.

Terzaghi, K., and R.B. Peck, 1967, Soil mechanics inengineering practice (2d ed.): New York, J. Wiley &Sons, Inc., 729 p.

Waples, D.W., and H. Kamata, 1993, Modelling poros-ity reduction as a series of chemical and physicalprocesses, in A.G. Doré et al., eds., Basin modelling:advances and applications: Norwegian PetroleumSociety (NPF) Special Publication 3, p. 303–320.

Wernicke, B., 1985, Uniform-sense simple shear of thecontinental lithosphere: Canadian Journal of Science,v. 22, p. 108–125.

Burial History and Porosity Evolution of Brazilian Upper Jurassic to Tertiary Sandstone Reservoirs 89

Page 106: Reservoir Quality Prediction in Sand and Carbonates

91

Chapter 7

A Geological Approach to PermeabilityPrediction in Clastic Reservoirs

Jonathan EvansBP Exploration

Poole, Dorset, England, United Kingdom

Chris CadeBP Norge

Stavanger, Norway

Steven BryantENIRICERCHE

Milan, Italy1

ABSTRACT

Permeability is a key parameter in determining the economic value of ahydrocarbon accumulation; however, our ability to predict the magnitude andrange of permeability in undrilled areas is poor. Traditional methods of perme-ability prediction are empirical and rely on developing relationships betweenpermeability and other parameters that may be predicted with greater confi-dence, such as porosity or lithology. These empirical methods may work wellin areas where there is sufficient calibration data, but extrapolation away fromwell data is prone to large errors (often by orders of magnitude).

An alternative approach to permeability prediction is to model the effectof geological processes such as burial and cementation on the pore structureof the rock and, hence, calculate the change in permeability. Through under-standing the effect of various geological processes on permeability, it is thenpossible to predict permeability from geological models. This approach hasapplications in both data-rich and undrilled areas.

The quantitative insight into which factors affect the permeability has beenprovided by computer modeling, which allows us to focus in on the mostimportant controls, such as grain size and the amount of cement or ductilegrains. Our ability to predict permeability in undrilled areas is now moreoften hampered by our inability to predict the variations in these controllingfactors rather than by any lack of understanding of permeability itself.

1Present address: Center for Subsurface Modeling, Texas Institute for Computational and Applied Mathematics, University of Texas at Austin, U.S.A.

Evans, J., C. Cade, and S. Bryant, 1997, A geologicalapproach to permeability prediction in clastic reser-voirs, in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones and car-bonates: AAPG Memoir 69, p. 91–101.

Page 107: Reservoir Quality Prediction in Sand and Carbonates

92 Evans et al.

INTRODUCTION

Porosity and permeability are important parametersthat help to define the commercial viability of an oil orgas accumulation. In particular, reservoir permeabilityis an important control on the flow rates that may beachieved from a well. As a result, the ability to predictpermeability has important commercial significance.

Permeability measures the ability of a rock to allowfluids to move through its pore system. It is a key fac-tor with respect to producing fluids from a reservoir.The controls on porosity are well understood, andmethods of porosity estimation are becoming wellestablished. In comparison, our understanding of thefactors controlling permeability is less advanced. Thischapter reviews the various methods that have beenused to help minimize the uncertainty inherent in per-meability prediction.

The data available for permeability prediction varywith the stage of a reservoir evaluation. At the wildcatstage, an assessment of permeability before drilling isessential to constrain the potential economic return. Thisusually will be based on regional porosity-permeability-depth trends together with sedimentological informa-tion; some burial history data may also be added. Inappraisal and development, a detailed description ofpermeability is required. Direct measurements of reser-voir characteristics from seismic reflection data, wirelinelogs, well tests, and core samples will be available. Pre-diction during these stages involves integration of per-meability measurements with information on reservoirsedimentology, together with seismic reflection andwireline log data, to fill the gaps between wells and pro-duce an overall reservoir description.

DETERMINATION OF PERMEABILITY

Permeability is the intrinsic characteristic of a mater-ial that determines how easily a fluid can pass throughit. In the petroleum industry, the darcy is the standardunit of permeability, but millidarcys (1 md = 10–3 dar-cys) are commonly used. Permeabilities in clastic reser-voir rocks may range from <0.1 md to >10 darcys. Thisintrinsic rock property is called absolute permeabilitywhen the rock is 100% saturated with one fluid phase.

The three main permeability measurement tech-niques are well testing, wireline tool analysis, and lab-oratory analysis of core samples.

Well Testing

Well testing can take various forms, but all involvethe measurement of a flow rate for fluid moving into thewell bore from the reservoir. The simplest test is a spin-ner survey, in which a turbine is moved up the well boreto record the location and velocity of any flow. Otherforms of testing, such as the drill-stem test, involve tak-ing measurements of pressure changes through timeeither before or after a restriction to flow. When thesepressure data are combined with measurements ofreservoir thickness, permeability can be calculated.

Well testing provides an average measurement ofpermeability across a certain reservoir interval. For oilor gas flows, well tests usually measure relative per-meability, rather than absolute permeability, sincemore than one fluid phase is present.

Wireline Measurements

Many methods have been proposed for obtainingpermeability measurements from wireline tool mea-surements. These include: (1) pressure/time measure-ment of formation fluids with the repeat formation testtool; (2) empirical correlation of permeability (fromcore analysis) with porosity and intergranular surfacearea (measured by wireline tools); (3) measurement ofmovable fluids with the nuclear magnetic resonancelog; and (4) correlation of permeability with Stoneleywave velocity measured by acoustic logging tools.

The applications of these methods have beenreviewed by Ahmed et al. (1991). Most of the methodsare at best qualitative—capable of distinguishing high-and low-permeability zones. The exceptions are for-mation test measurements and the standard core-derived permeability vs. porosity regression method.The latter is valid only for formations similar to thecalibrated formation.

Core Analysis

Core analysis allows direct measurement of poros-ity and permeability under controlled laboratory con-ditions. Measurements can be made at three scales:rotary sidewall core [samples <2.5 cm (1 in.) long],core plugs [samples 2.5–4 cm (1–1.5 in.) long], andwhole core [samples ≤60 cm (2 ft) long].

Such measurements give an accurate representationof a particular core sample under specific laboratoryconditions. Extrapolation to field conditions must bedone with care.

Routine core analysis is normally carried out on coreplugs taken every 30 cm (1 ft) through whole core. Thisprovides data on porosity and air permeability (Ka). Acorrection is usually applied to the Ka values to giveequivalent liquid permeability (KL) (Klinkenberg, 1941).

Special core analysis (SCAL) may be performed ona selection of plugs from the reservoir interval todetermine brine permeability (Kb). This may be mea-sured over a range of confining pressures to determineKb at overburden pressure. Other SCAL methods candetermine a range of petrophysical parameters ifrequired (e.g., capillary pressure, relative permeabil-ity, and formation factor).

CONTROLS ON PERMEABILITY

In clastic rocks, permeability is determined by thesize of the pore throats present in the rocks and by thenumber of connected pores. Permeability predictioninvolves understanding how various geological factorsaffect these fundamental controls. In unconsolidatedsands, the important factors are the grain size and sort-ing (Beard and Weyl, 1973). Rocks with coarser grain

Page 108: Reservoir Quality Prediction in Sand and Carbonates

sizes will tend to have larger pore throats and, therefore,higher permeability. Rocks with poorer sorting willhave smaller mean pore-throat diameters and, therefore,lower permeability than better sorted rocks with thesame mean grain size. The presence of detrital clays willlead to smaller pore throats and less-connected pores,which reduces permeability. During burial, compactionreduces the size of pore throats and eventually blocksthem off entirely, so again permeability is reduced. Therate of compaction and the rate of pore-throat blockingdepends on the proportion of ductiles present; this alsoaffects the permeability (Gluyas and Cade, this volume).The precipitation of cements similarly reduces the sizeand number of pore throats; different cement styles willreduce the permeability at different rates. Ethier andKing (1991) illustrate a general understanding of thesecontrols (Figure 1), but with little or no quantitativedetail, the value of such trends is limited.

EMPIRICAL APPROACHES TOPERMEABILITY PREDICTION

Empirical techniques use a calibration data set (e.g.,data from core samples) and multiple regressionanalysis to determine the relationship between rockproperty variables and reservoir quality. The cali-brated regression relationships are then used to pre-dict reservoir quality in different settings but withinthe range of the variables comprising the calibrationdata set. Dutton and Diggs (1992) and Bloch (1991)describe the most frequently used application of thisapproach, in which relationships between measuredporosity and permeability (usually ambient heliumporosity and single-phase gas permeability), and tex-tural and mineralogical variables (usually measuredon thin sections), are investigated. Commonly usedvariables are grain size, sorting, matrix clay content,volume of individual cements, total cement volume,and point-counted interparticle porosity.

A variation on the empirical approach is describedby Ehrlich et al. (1991). Using the observation that, evenin a single formation, permeability commonly varies byseveral orders of magnitude, they conclude that theconfiguration, rather than the absolute value, of poros-ity is the control on permeability. To characterize thepore system configuration, Ehrlich et al. (1991) makemeasurements of pores in two dimensions (on polishedthin sections) and combine these with pore-throat sizedistribution data (from mercury porosimetry) todevelop a simple pore system model. For selected datasets, a good relationship between the simple pore sys-tem model and measured permeability has been estab-lished. It is unlikely, however, that we would be able topredict confidently the pore type and pore-throat sizedistribution parameters in an undrilled sandstone.

In predicting permeability ahead of drilling, the crite-ria for success of any method must be that it establishesa quantitative link between measured permeability andanother (or several other) rock parameter(s), and thatthose correlative parameters can themselves be pre-dicted from a geological model. Many empiricalapproaches fail the second of these criteria.

APPLICATIONS OF THE EMPIRICALAPPROACH

In areas with sufficient well data (either core analy-sis, log, or well test data) to define significant regionalor field porosity–permeability and porosity–depthregressions, the empirical approach described abovecan often be successfully used to predict porosity andpermeability in areas away from well control.

This method is the one most commonly used inmature provinces, and gives good results providedthere is not too much scatter in the data. However, thescatter is often such that the uncertainty in permeabil-ity prediction may cover several orders of magnitude.

Some of this scatter may be due to textural varia-tion, controlled in turn by sedimentary facies andlithology. If sedimentological information is available(from core logs or reservoir models), lithofacies can betaken into account by plotting the poroperm values foreach lithofacies separately. Often the regression rela-tionships for a given lithofacies will be better thanthose for the whole data set, since variations in grainsize, clay content, and so forth will be reduced. Thecombination of empirical relationships for each facieswith a sedimentological reservoir model may thenproduce a reasonable description of reservoir perme-ability variations.

Further insight may be obtained through includingmineralogical (e.g., from modal point counting), tex-tural (e.g., grain size, sorting), and SCAL (e.g., criticalpore-throat size, Kb) data in the regression analysis. Inmany cases, a few parameters will explain most of thevariation in permeability (e.g., grain size, sorting, lithiccontent).

Example 1—Permeability Prediction While Drilling

Hogg et al. (1996) have recently presented a novelapplication of the empirical approach to permeabilityestimation. In the Triassic fluvial Sherwood Sandstonereservoir of the Wytch Farm field (onshore U.K.), per-meability is controlled principally by porosity andgrain size. Using previously drilled wells in the field,an empirical correlation was determined betweenporosity and permeability for a range of grain-sizeclasses (Figure 2).

By measuring porosity (from logging-while-drillingdensity tool) and grain size (from sieve analysis of cut-tings) during drilling, it is possible to estimate the per-meability of the sandstone while drilling. This methodwas applied during drilling of extended-reach wells inorder to ensure that a certain minimum productivityindex (PI) is achieved before the well is stopped.

Test results from several extended-reach wells showthat PI can be predicted with a high degree of accuracyusing this technique (Figure 3). The main uncertainty isthe density of the saturating fluid, which must beassumed when calculating porosity. An additionalbenefit is that the predicted permeability–depth plotcan be used to optimize perforation intervals.

A Geological Approach to Permeability Prediction in Clastic Reservoirs 93

Page 109: Reservoir Quality Prediction in Sand and Carbonates

94 Evans et al.

Example 2—Mapping Permeability inClyde Field, UKCS

The Clyde Field [United Kingdom ContinentalShelf (UKCS)] contains oil in Upper Jurassic shallowmarine sands. The field has been on production forsome time, and infill drilling is currently in progress.

A study of the controls on porosity and permeabilityin the Clyde Field reservoir sandstone was performedin order to reduce the uncertainty associated with per-meability prediction ahead of drilling infill wells and toimprove the mapping of permeability between wells(J. Gluyas, 1995, personal communication).

Accurate prediction of permeability within theClyde Field is difficult because in some reservoir zonespermeability varies by up to 4 orders of magnitude fora given porosity. Porosity varies little within andbetween reservoir zones. Porosity and permeabilitydisplay no correlation. There is, however, a good cor-relation between the maximum grain size and mea-sured permeability from the core (Figure 4). For eachreservoir layer that displays a grain size variation,there is a systematic and predictable trend across thefield (e.g., Figure 5). Thus, the correlation betweengrain size and permeability can be used to predict per-meability ahead of drilling or to map permeability inuncored areas for reservoir simulation purposes.

Despite the success of these empirical methodswhen dealing with a particular field, correlative tech-niques are limited by the need for pre-existing data.Also, since no insight is gained into the processes con-trolling permeability, there is no basis for extendingpredictions beyond the range of calibration data.Therefore, permeability prediction in areas with littleor no well data requires another approach.

MODELING APPROACHES TOPERMEABILITY PREDICTION

The major factors controlling sandstone permeabilityare grain size, sorting, compaction, and cementation(Cade et al., 1994). Computer models may be used tohelp understand the effects of these parameters on per-meability. The effects of compaction (vertical shorteningof a rock volume) and cementation (various types ofpore filling) have been modeled using a numerical rep-resentation of a real porous medium, a sphere-pack oflike-sized, randomly packed grains (Bryant et al., 1993a,b; Cade et al., 1994). Using this approach, the porosity-permeability trends that result from the progressiveapplication of various diagenetic processes, either ontheir own or in combination, can be understood. Theeffects of grain size on the porosity–permeability trends

INCREASINGGRAIN SIZE

INCREASINGVISIBLEPOROSITY

INCREASIN

G RELI

EF

INCREASIN

G

CEMENTATIO

N/COM

PACTION

IMPROVED SORTING

INCREASING MICROPOROSITY

INC

RE

AS

ING

INTE

RG

RA

NU

LAR

C

LAY

POROSITY

PE

RM

EA

BIL

ITY

Figure 1. An interpretation ofthe effects of various controlson porosity and permeability(after Ethier and King, 1991).

Page 110: Reservoir Quality Prediction in Sand and Carbonates

can also be modeled. The extension of this modeling toaccount for less than perfect sorting remains problem-atic (Panda and Lake, 1994, 1995). Clearly, sorting is offundamental importance, but at present there is only anempirically based correction for sorting variation (Beardand Weyl, 1973). Despite this drawback, computer mod-eling will, in many cases, provide the basis for enhancedpredictions of permeability in combination with predic-tions of texture, diagenetic modification, and porosity.

The “process-oriented” approach described byBloch (1991) and Bloch and Helmold (1995) focuseson modeling diagenetic processes in an undrilled

area, based on chemical and mathematical models,and the effects of those processes on reservoir qual-ity. There are two important limitations to thisapproach: first, there is the uncertainty associatedwith the subsurface geological model, and how itimpacts on the thermodynamics and kinetics of themodels; second, there is the lack of a detailed quanti-tative understanding of how diagenetic processescontrol permeability. To date, the quantification ofthe impact of specific controls, particularly diageneticcontrols, has been either formation specific (andtherefore not widely applicable) or very general.

A Geological Approach to Permeability Prediction in Clastic Reservoirs 95

Figure 2. Variation in permeability with horizontal core plug porosity and grain size class in the SherwoodSandstone Formation for three wells in the Wytch Farm field, onshore UK (after Hogg et al., 1996).

Page 111: Reservoir Quality Prediction in Sand and Carbonates

96 Evans et al.

APPLICATIONS OF THE MODELINGAPPROACH

Example 4—Frontier Exploration

Economic success in many frontier hydrocarbonbasins is dependent on high oil production rates. Pro-duction rate is in turn controlled by the permeability ofthe reservoir rocks. In basins where potential reservoirtargets are deep, reservoir quality is a key risk to dis-covering commercial hydrocarbon volumes. Conven-tionally, reservoir quality would be predicted based on

comparison with analog basins or by calibration withnearby wells (see earlier discussion of empiricalapproaches). In frontier basins, however, few data maybe available to help assess the depth to economic base-ment. In such cases, where it may be unclear whichanalogs are appropriate, a process-oriented approachto estimating the depth limit of effective permeabilitymay be more appropriate.

Described below is a method to estimate the depth toeconomic basement that is based on geological models ofporosity–depth relationships and porosity–permeability

Figure 3. Predicted permeability–depth profiles and productivity index (PI) for well L98/6-F19 in the Wytch Farmfield, onshore UK. Actual well trajectory and test results are shown for comparison (after Hogg et al., 1996).mTVDBRT = meters true vertical depth below rotary table; OWC = oil-water contact; brt = below rotary table;PI = productivity index; sbb/d/psi = stock tank barrels per day per psi.

Page 112: Reservoir Quality Prediction in Sand and Carbonates

relationships (Bryant et al., 1993a; Gluyas and Cade, thisvolume). The method is illustrated by reference to a realexample taken from a frontier basin in Southeast Asia.

To help in ranking prospects with different depthsand containing sands with potentially different com-positional maturity, “cutoff depths” below whichreservoir quality will be insufficient to provide eco-nomic flow rates must be defined. Cutoffs have beencalculated for a range of possible geological scenarios.The results provide a guide to what geological condi-tions will allow a given prospect to be economic. Byassessing the chance of the necessary conditions occur-ring, the risk on reservoir quality can be estimated.

The likely range of geological variation was takenfrom the best and worst cases seen in nearby wells.Best case: Uncemented, coarse-grained pure quartzsand. Worst case: Very fine grained sand containing25% ductile lithic grains and cemented by 20% quartzcement. Based on required flow rates and likely reser-voir thickness, economic permeability cutoffs weregiven as 18 md for gas and 45 md for oil at surface con-ditions. At reservoir conditions, these are estimated toreduce by a factor of 3, because of overburden pres-sure and relative permeability effects.

The general method of calculating depth cutoffs isstraightforward. Using the economic permeability cutoffas a starting point, we must calculate what the porosityof the rock was after it had been compacted but beforeany cements started to form (this porosity value isknown as the compactional porosity and gives an upperlimit to the porosity in normally pressured rocks). Usingthis porosity, the cutoff depth is calculated from com-paction curves.

For uncemented rocks, the reservoir porosity (or “cut-off porosity”) is equal to the compactional porosity.However, for rocks containing cement, the compactional

porosity is calculated by adding the volume of cement tothe cutoff porosity.

The depth cutoffs were calculated using the followingapproach:

1. Construct the permeability-porosity curve for therelevant grain size, sediment composition (i.e.,clean or ductile-bearing), and cement type.

2. Using the appropriate permeability cutoff (18 or45 md), read the equivalent porosity (Figure 6).

3. Using the compactional porosity obtained above,calculate the cutoff depth using appropriate com-paction curves (Figure 7).

The results of cutoff depth calculations for two mostlikely geological cases are presented in Tables 1 and 2.The values shown in the cutoff tables assume that thereservoirs are normally pressured. The cutoff depthswill increase by ~550 m for every 1000 psi of overpres-sure (assuming overpressure was present beforecements formed).

Prior to drilling prospects in this basin, reservoireffectiveness was perceived to carry a high risk,because ductile-rich sands were considered likely.Temperature gradients are such that quartz cementsare likely to be present at depths greater than ~3500 m.

Since these original predictions were made, severalwells have penetrated these potential reservoirs. Thesandstones have proved to be ductile rich. The porosi-ties are consequently low (15%–20% at 3000 m) and the

A Geological Approach to Permeability Prediction in Clastic Reservoirs 97

1

10

100

1000

10,000

0 1 2 3 4 5 6

Co

re p

erm

eab

ility

(m

d)

Maximum grain size (phi units)

Coarse Medium Fine V.Fine Silt

Figure 4. Correlation between maximum grain sizein phi units (–log2 grain size in millimeters) andarithmetic mean core permeability for layer AL3 inClyde Field, offshore UKCS.

Figure 5. Variation of grain size within layer AL3across Clyde Field, offshore UKCS. Grain sizevaries between medium sand in the south of thefield to silt in the north. The grain size variationscan be used to map permeability variations in thiscase. f = fine; m = medium; vf = very fine.

CLYDE FIELD

m. sand

f. sand

vf. sand

silt

Page 113: Reservoir Quality Prediction in Sand and Carbonates

98 Evans et al.

permeabilities are subeconomic below 3000–3500 m(Worden et al., 1996).

Example 5—Appraisal of Unconsolidated Reservoirs

Unconsolidated reservoirs are common in manyparts of the world (e.g., Gulf of Mexico, North Sea, andWest Africa). There are particular problems in deter-mining the permeability of such reservoirs. Cores aredifficult to obtain, and core analysis results are oftenunreliable due to “repacking” of grains. Also, dynamictesting is difficult, because the formations may collapseduring flow stimulation. However, unconsolidatedreservoirs are usually uncemented, so porosity andpermeability are largely controlled by compaction.

Permeability may be accurately predicted if porosityand grain size are known or may be predicted.

The Harding Field (offshore UKCS) has an uncon-solidated sandstone reservoir. During early appraisal,a knowledge of reservoir permeability was critical todetermining the economic viability of the field. Twoconflicting permeability estimates were available fromdifferent sources. Well-test data suggested a perme-ability of ~9–10 darcys, whereas core analysis dataimplied lower values of ~3–4 darcys. The choice ofpermeability value had implications for the time towater breakthrough and the optimum height in thereservoir at which to drill a horizontal well.

Using grain size measurements from sieve analysisof core samples (Table 3), together with porosity

10 Porosity Units

10 Porosity Units1 2 3 4

Clean Sands - Porosity Reduction by :Compaction (1,2),Compaction, then 10% Quartz Cement (3,4)

Oil Cutoff

Gas Cutoff

Per

mea

bilit

y (m

d)

Porosity (%)

100,000

10,000

1000

100

10

0 5 10 15 20 25 30 35

Figure 6. Example curve forcalculating porosity cutoffsfor a given permeability inclean sands in which porosi-ty reduction is by com-paction and/or quartzcementation. In this case,permeability cutoffs of 18and 45 md (gas and oil) areequivalent to porosity cut-offs of 6% and 7.5%. In thecase where the rocks containquartz cement, the amount ofcement (here 10 vol%) mustbe added to the cutoff poros-ity in order to determine the“compactional porosity,”which is used to derive thedepth cutoff for a given per-meability (Figure 7).

Figure 7. Example of how todetermine depth cutoffs fora given permeability. Usingthe porosities determined inFigure 6, the equivalent cut-off depth may be found byreference to a clean sandcompaction curve (Gluyasand Cade, this volume).

Page 114: Reservoir Quality Prediction in Sand and Carbonates

derived from logs (estimated to be 34%–35%), the pre-dicted porosity-permeability relationships of Bryant etal. (1993a) were used to estimate the likely permeabilityof the reservoir (Figure 8). This was then corrected toreservoir conditions (an overburden correction of ~1.5was applied) to allow comparison with the test data.The permeability estimated in this way was 9–12 darcys(range due to grain size variations), which was muchcloser to the test results than to the core analysis data.The test data were therefore used for reservoir simula-tion, and the core analysis results were discarded.

Subsequently, it was discovered that the core analy-sis results had been in error. The analytical methodhad not been appropriate for such high-permeabilitysamples, so that, in effect, the permeability of the testapparatus itself had been measured. More carefulmeasurements were made, which confirmed the per-meability modeling and well-test data.

CONCLUSIONS

Through recognizing the important controls onpermeability in clastic rocks, namely, grain size, sort-ing, compaction, and cementation, it is possible topredict permeability in undrilled areas by application

of geological models. In areas where there are exist-ing core data, empirical methods that relate perme-ability to other predictable parameters (e.g., grainsize variation) will give good results. However, inareas away from well control or in fields where thecontrols on permeability are complex, predictionsbased on geological models combined with perme-ability modeling results are likely to give betterresults. A combination of empirical and geologicalmodeling approaches will often give the best results,even in areas where there is abundant data (seeexamples in Cade et al., 1994). The quantitativeinsight into the way in which different factors affectpermeability, which has been provided by computermodeling (Bryant et al., 1993a; Panda and Lake,1995), allows us to focus on the most important con-trolling factors. These are often grain size and theamount of cement or ductile grains. Our ability topredict permeability in undrilled areas is now moreoften hampered by our inability to predict these con-trols rather than by any lack of understanding of per-meability itself. This shifts the emphasis back to thesedimentologists and geologists to better constraintheir geological models so that the uncertainty in thepossible range of permeabilities may be reduced.

A Geological Approach to Permeability Prediction in Clastic Reservoirs 99

Table 1. Case 1: Pure Quartz Sand, 10% Quartz Cement Porosity Reduction byCompaction, Then Quartz Cement.*

Grain Size, Cutoff Cutoff CutoffSorting Permeability (KL) Porosity (%) Depth (m)

vfs, mod. sorted 18 md 12 300045 md 15 2200

fs, mod. sorted 18 md 7.5 >500045 md 9.5 >5000

ms, mod. sorted 18 md 5 >500045 md 6.5 >5000

cs, mod. sorted 18 md 3.5 >500045 md 4.5 >5000

*cs = coarse sand; fs = fine sand; ms = medium sand; vfs = very fine sand.

Table 2. Case 2: Sand Containing 25% Ductiles, No Cement, PorosityReduction Solely by Compaction.*

Grain Size, Cutoff Cutoff CutoffSorting Permeability (KL) Porosity (%) Depth (m)

vfs, mod. sorted 18 md 19.5 265045 md 21 2350

fs, mod. sorted 18 md 17 320045 md 18 2950

ms, mod. sorted 18 md 15.5 360045 md 16.5 3350

cs, mod. sorted 18 md 14.5 395045 md 15 3750

*cs = coarse sand; fs = fine sand; ms = medium sand; vfs = very fine sand.

Page 115: Reservoir Quality Prediction in Sand and Carbonates

100 Evans et al.

ACKNOWLEDGMENTS

Thanks to BP Exploration for permission to publishthese data and ideas. Our understanding of permeabil-ity prediction has evolved over the past five yearsthrough interaction and discussions with many pastand present colleagues. In particular, we would like tothank Craig Smalley, Tony Mitchell, AndrewBrayshaw, Sue Raikes, Dave Mellor, Harit Trivedi, EdWarren, Andrew Hogg, Tim Primmer, Shona Grant,Jon Gluyas, Norman Oxtoby, Richard Worden, KevinSchofield, and Mike Bowman.

REFERENCES CITEDAhmed, U., S.F. Crary, and G.R. Coates, 1991, Perme-

ability estimation: the various sources and theirinterrelationships: Journal of Petroleum Technology,v. 42, p. 578–587.

Beard, D.C., and P.K. Weyl, 1973, Influence of textureon porosity and permeability of unconsolidatedsand: AAPG Bulletin, v. 57, p. 349–369.

Bloch, S., 1991, Empirical prediction of porosity andpermeability in sandstones: AAPG Bulletin, v. 75, p. 1145–1160.

Bloch, S., and K.P. Helmold, 1995, Approaches to pre-dicting reservoir quality in sandstones: AAPG Bul-letin, v. 79, p. 97–115.

Bryant, S.L., C.A. Cade, and D.W. Mellor, 1993a, Per-meability prediction from geological models:AAPG Bulletin, v. 77, p. 1338–1350.

Bryant, S.L., D.W. Mellor, and C.A. Cade, 1993b, Phys-ically representative network models of transport inporous media: American Institute of ChemicalEngineers Journal, v. 39, p. 387–396.

Cade, C.A., J. Evans, and S.L. Bryant, 1994, Analysis ofpermeability controls: a new approach: Clay Miner-als, v. 29, p. 491–501.

Dutton, S.P., and T.N. Diggs, 1992, Evolution of poros-ity and permeability in the Lower CretaceousTravis Peak Formation, East Texas: AAPG Bulletin,v. 76, p. 252–269.

Ehrlich, R., E.L. Etris, D. Brumfield, L.P. Yuan, and S.J.Crabtree, 1991, Petrography and reservoir physicsIII: physical models for permeability and formationfactor: AAPG Bulletin, v. 75, p. 1579–1592.

Ethier, V.G., and H.R. King, 1991, Reservoir qualityevaluation from visual attributes on rock surfaces:methods of estimation and classification from drillcuttings or cores: Bulletin of Canadian PetroleumGeology, v. 39, p. 233–251.

Gluyas, J.G., and C.A. Cade, this volume, Prediction ofporosity in compacted sands, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir 69,p. 19–28.

Hogg, A.J.C., A.W. Mitchell, and S. Young, 1996, Pre-dicting well productivity from grain-size analysisand logging while drilling: Petroleum Geoscience,v. 2, p. 1–15.

Table 3. Grain Size Data Determined by Sieve Analysis of UnconsolidatedCore Samples.

9/23b-7 9/23b-8Sieve Data (Harding Central) (Harding South)

Mean grain size (µm) 243 240Median grain size (µm) 243 238Sorting mod. well or well mod., mod. well, or well

1

10

100

1000

10,000

100,000

0 5 10 15 20 25 30 35

Modeled Permeability-Porosity Trend

Well-Test Permeability

Core Analysis Permeability

Per

mea

bilit

y (m

d)

Porosity (%)

Figure 8. Modeled porosity–permeability curves forsands in the Harding Field(offshore UKCS) based onsieve data. Well-test data(corrected to surface condi-tions) and routine coreanalysis data are shown forcomparison. The modelingresults confirm that the testdata are valid.

Page 116: Reservoir Quality Prediction in Sand and Carbonates

Klinkenberg, L.J., 1941, The permeability of porousmedia to liquids and gases: Drilling and ProductionPractices, API, Dallas.

Panda, M.N., and L.W. Lake, 1994, Estimation of single-phase permeability from parameters of parti-cle size distribution: AAPG Bulletin, v. 78, p. 1028–1039.

Panda, M.N., and L.W. Lake, 1995, A physical modelof cementation and its effects on single-phase per-meability: AAPG Bulletin, v. 79, p. 431–443.

Worden, R., M. Mayall, and J. Evans, 1996, The effect oflithic grains on porosity and permeability in Tertiaryclastics, South China Sea: Journal of the GeologicalSociety of London.

A Geological Approach to Permeability Prediction in Clastic Reservoirs 101

Page 117: Reservoir Quality Prediction in Sand and Carbonates

103

Chapter 8

Detecting Permeability Gradients inSandstone Complexes—Quantifying the

Effect of Diagenesis on FabricRobert Ehrlich

Department of Geological Sciences, University of South CarolinaColumbia, South Carolina, U.S.A.

Mark C. BowersConoco Incorporated

Houston, Texas, U.S.A.

Virginia L. RiggertAmoco Production Co.

Denver, Colorado, U.S.A.

Christopher M. PrinceDepartment of Geological Sciences, University of South Carolina

Columbia, South Carolina, U.S.A.

ABSTRACT

Matrix permeability, the permeability associated with measurements onsmall samples, is controlled by depositional fabric and diagenesis. Predictionof matrix permeability requires: (1) specification of a fabric, (2) specificationof the diagenetic state, and (3) a means to assess both factors in a sample settaken from a target basin. The data from the sample set can be used toextrapolate or interpolate within the basin or may be used to calibrate fabricresponse to basin history data (e.g., thermal history). The effects of fabricand diagenesis on the sample set can be determined using a combination ofimage analysis data and mercury porosimetry data.

Strong correlations exist between permeability and grain size of unconsoli-dated sands and gravels, with permeability increasing exponentially withincreasing grain size. Permeability in clastic fabrics is controlled by networksof packing flaws, characterized by large pores connected by large porethroats. Such circuits comprise only a fraction of the porosity and representthe effective flow component of porosity. Diagenesis usually brings aboutpermeability reduction, but preferentially affects the grains in close-packedarrangements that separate the networks of packing flaws. A methodology

Ehrlich, R., M.C. Bowers, V.L. Riggert, and C.M.Prince, 1997, Detecting permeability gradients insandstone complexes—quantifying the effect ofdiagenesis on fabric, in J.A. Kupecz, J. Gluyas, andS. Bloch, eds., Reservoir quality prediction in sand-stones and carbonates: AAPG Memoir 69,p. 103–114.

Page 118: Reservoir Quality Prediction in Sand and Carbonates

104 Ehrlich et al.

INTRODUCTION

Reservoir-scale permeability prediction requiresknowledge of many properties at many scales and,therefore, requires a multidisciplinary team thatincludes petrologists, stratigraphers, basin modelers,and others. Because reservoir-scale permeability pre-dictions are ultimately derived from matrix permeabil-ity measurements (usually associated with smallvolumes of rock), it is important to understand howmatrix permeability varies within a basin. The objec-tive of this chapter is to discuss how matrix permeabil-ity prediction can be used as a lead-in to reservoir-scalepermeability prediction. We discuss methods for pre-diction of the highest permeability possible that mightbe encountered as a function of basin location, given aparticular fabric, because only a small fraction of theporosity of a sandstone contributes to permeability(Ehrlich et al., 1991b; Prince et al., 1995). We show howthe porosity components that most influence perme-ability can be identified, and how the rate of change ofthe size of associated pore throats with respect to depth(or basin location) can then be determined.

Prediction of matrix permeability (hereafter referredto simply as permeability) must take into account boththe depositional fabric and the diagenetic modificationof that fabric. Most of the permeability contrastsobserved in a single core arise from grain size variation,because all samples have, to a first approximation, acommon postdepositional history resulting in a com-mon diagenetic state. Increasing diagenesis alters orobscures the relationship between depositional fabricand permeability, but it never completely erases it. Sig-nificant permeability contrasts observed in a single coreare commonly associated with grain size variationbecause all samples have a common postdepositionalhistory resulting in a common diagenetic state. Givenrock of similar composition, diagenesis varies spatiallyas a response to gradients in pressure, temperature, andfluid chemistry. That is, individual components of dia-genesis exist in the form of a diagenetic gradient; in the-ory, this component can be mapped over a basin. In ourexperience, such gradients are common in sandstoneswhere reduction in permeability is largely due to fac-tors such as quartz overgrowth development and/orcompaction and pressure solution.

Intergranular pore shape is affected by the shapes ofthe bounding sand grains, as well as whether they arelocated in close-packed or loose-packed domains.Intergranular pore size is a function of grain size, mod-ified according to whether the pore exists in close-packed (relatively smaller) or loose-packed (relatively

larger) domains. Pore types can be defined as a popula-tion of pores with a characteristic size and shape. Anobjective quantitative porosity classification into poretypes that is rapid and precise can be achieved by usingimage analysis procedures described in Ehrlich et al.(1991b). The automated classification is consistent withconventional classification, while easily capturing dif-ferences in size, shape, and type of porosity (intergran-ular, intragranular, and moldic). Image analysis breaksdown porosity complexes into as many pore types asdemanded by variations in depositional fabric (includ-ing grain size) or differential effects of diagenesis.Reservoirs commonly contain four to seven pore types,depending on grain size variability (Horkowitz, 1987;Ehrlich et al., 1991b; Bowers, 1992; Murray et al., 1994;Riggert, 1994), whereas individual samples are usuallydominated by one or two pore types.

Permeability is strongly dependent on the effi-ciency of the intergranular porosity, of which moldicand intragranular porosity generally contribute little.Intergranular porosity covers a wide range of sub-types; the quantitative objective characterization ofthese subtypes is crucial to permeability prediction.Intergranular porosity falls into two types: that foundin close-packed domains and that found in loose-packed domains (Graton and Fraser, 1935; Prince etal., 1995). These types can be expressed in a variety ofways, depending on grain size and sorting. Loose-packed porosity (“packing flaws”) has large-scale spa-tial continuity and is associated with larger pore throatsizes. These two factors make loose-packed domainsthe major contributors to permeability (McCreesh etal., 1991; Anguy et al., 1994; Prince et al., 1995).

THE EFFECTIVE COMPONENT OF POROSITY

Permeability is independent of porosity in unconsol-idated sands. Under progressive diagenesis, a roughrelationship between porosity and permeability mayarise. However, at the median porosity in a sample set,permeability commonly varies by more than 2 ordersof magnitude. Among samples with the same perme-ability, porosity may vary by more than 10 porosityunits (e.g., 10%–20%) (Figures 1A, 2A). Much of thisscatter is due to the effect of variations in grain size,with porosity being preferentially reduced in finergrained sandstones.

Subsets of porosity are much more highly correlatedwith permeability than core (bulk) porosity, suggestingthat some parts of the pore system do not support much

has been developed over the past decade that quantifies thin-section–baseddata precisely enough to estimate the effects of grain size and diagenesis onthe rock fabric with respect to flow properties. Such rock physics data arenecessary for permeability prediction as a function of basin position.

Page 119: Reservoir Quality Prediction in Sand and Carbonates

Detecting Permeability Gradients in Sandstone Complexes––Quantifying the Effect of Diagenesis on Fabric 105

Figure 1. Relationships betweenpermeability and subsets ofporosity for Miocene sand-stones, Gulf of Thailand: (A)core porosity, (B) total opticalporosity (TOP), (C) TOP portionconsisting of PT4 (pore type 4)and PT5, and (D) number ofpores per unit area of PT4 andPT5. Squares represent PT4 anddots represent PT5.

(A)

(B)

(C)

(D)

Page 120: Reservoir Quality Prediction in Sand and Carbonates

106 Ehrlich et al.

flow, even at the scale of matrix permeability. This canbe illustrated by observing the increase in correlationbetween permeability and successive refinements ofporosity (Figures 1, 2), using image analysis of petro-graphic thin sections. Digital image analysis at lowmagnification (<100×) will commonly yield a value ofporosity [total optical porosity (TOP)] less than themeasured value because small pores and small-scaleroughness on pore walls cannot be resolved. Total opti-cal porosity, however, always correlates more highlywith permeability than does core porosity (Figures 1B,2B); TOP can be subdivided into portions associatedwith each pore type. The amount of TOP associatedwith one or more pore types (the product of the relativeproportion of a pore type and TOP) is more highly cor-related with permeability (Figures 1C, 2C) than is eitherTOP or core porosity. The number of pores per unit areaof these pore types is also highly correlated with

permeability (Figures 1D, 2D). The high correlationbetween certain pore types and permeability impliesthat such pores must be connected by relatively largethroat sizes. The throat sizes associated with pores ofeach type can be quantified by relating the pore typedata with mercury-injection porosimetry data.

PORE TYPES AND THROAT SIZE

The amount of porosity lying behind pore throats ofvarious sizes can be determined from mercury-injectioncapillary pressure tests. McCreesh et al. (1991) foundthat, based on statistical analysis, different pore typestend to control different portions of the capillary pres-sure curves; that is, different pore types tend to fill in dif-ferent pressure ranges. This can occur only if pores oflike type are mutually adjacent, forming circuits charac-terized by a common throat size. Recently, Prince et al.

(A) (B)

(C) (D)

Figure 2. Relationships between permeability and subsets of porosity for Upper Carboniferoussandstones in Oklahoma: (A) core porosity, (B) total optical porosity (TOP), (C) TOP portionconsisting of PT3 and PT4, and (D) number of pores per unit area of PT3 and PT4.

Page 121: Reservoir Quality Prediction in Sand and Carbonates

(1995) optically resolved such circuits using filteredFourier transforms of large, high-resolution images; Rig-gert (1994) demonstrated the existence of these flow cir-cuits by analysis of suites of capillary pressure curves.

The pore type–throat size relationship is based on aset of regression analyses in which the pore type propor-tions are used to predict the amount of saturation in suc-cessive pressure intervals from the mercury-injectioncurves (McCreesh et al., 1991). One equation is producedfor each pressure interval that relates the pore type pro-portions to the mercury saturation. The set of equationsis then used to calculate the distribution of throat sizesand mean throat size associated with each pore type.

PORE TYPES AND PERMEABILITY

Ehrlich et al. (1991a) demonstrated how the associa-tion between pore type and throat size can be used tomodel permeability. Using a modified Hagen-Poiseuilleversion of Darcy’s Law, they showed that permeabilityis proportional to the product of the number of pores ofeach type per unit cross-sectional area and the fourthpower of the associated pore throat size. The model isbased on the assumption that the flow paths are rela-tively straight and parallel, and the number of effectivethroats is proportional to the number of pores. Porethroats at high angles to the pressure gradient are inef-fective. With this model, the amount of permeabilitycontributed by each pore type can be determined.

The modified Hagen-Poiseuille permeability modelis effective over a range from <1 md to several darcys.In sandstones with permeabilities >20 md, commonlyonly one or two pore types contribute most of the per-meability. The spatial rate of change of the throat sizeof such dominant pore types can be determined from areference set of cores, and that function can be used toprovide an estimate of the maximum permeability thatmay be encountered for such rocks.

GRAIN SIZE AND PERMEABILITY INLITHIFIED SANDSTONE

Pore typing automatically takes into account bothdepositional and diagenetic effects, because any changein grain size and any diagenetic event affects the sizeand/or shape of a pore. Grain size and packing are themajor depositional properties at perm plug scale. Therelationship between grain size and permeability is, toour knowledge, only documented for unconsolidatedsands (Shepard, 1989). Shepard stated that permeabilityis an exponential function of grain size, with the rela-tionship proportional to grain size raised to a powerranging from about 1.3 to 2 in. well-sorted sands. Lithi-fication commonly reduces the permeability at all grainsizes, but little is known concerning the relationshipbetween permeability in consolidated sandstones.

Because investigations in lithified sandstones havenot been published, some researchers have assumedthat the relationship remains exponential after lithifi-cation. One problem in verifying this assumption isthe determination of grain size in an indurated rock,

because three problems have had to be overcome: (1)many sandstones are not friable enough to disaggre-gate effectively; (2) acquisition of overgrowths biasdirect measurements on quartz grains observed inthin section; and (3) measurements of grain size inthin section are generally biased by the fact that agrain may not be cut by the plane of section near itsdiameter. Therefore, the distance from grain center tograin boundary represents an apparent grain size,because the distance of magnitude is influenced bythe location of the plane of section relative to thegrain “equator,” by overgrowth development, and bypressure solution. Prince et al. (1995) used a two-dimensional fast Fourier transform on thin-section–scale binary images to quantify the spatialfabric of sandstones. They pointed out that the cen-ter-to-center distances between pores are a goodapproximation of the distances between grain centersunaffected by the biases mentioned above.

A common assumption is that the grain size–permeability relationship observed in unconsolidatedsandstones by Shepard (1989) also holds true for lithi-fied sandstones. That is untrue, as shown by work doneby C.M. Prince et al. (personal communication) on asmall portion of their data for a Carboniferous sand-stone (Perry Sandstone) from the Cherokee Basin, Okla-homa. Grain sizes range from ~100 to 250 µm, withpermeabilities ranging from ~0.5 to 500 md. Resultsshown in Figure 3 indicate that extrapolations of thegrain size–permeability relationships in unconsolidatedsands should be used with caution when trying to char-acterize well-cemented sandstones in samples of lowpermeability (<10 md). This is especially true wherediagenesis has produced patchy fabrics; no grain size-permeability relationship exists (Figure 3). The PerrySandstone data also indicate the grain size–permeabil-ity relationship is exponential for samples with uni-formly altered fabrics (e.g., rimming cements, quartzovergrowths), with an exponent much greater than thatobserved in unconsolidated sands (5 in the consolidatedPerry Sandstones vs. 1.3–2.0 in unconsolidated sands)(Figure 3). The Perry Sandstone is similar to manyPaleozoic sandstones we have studied; it is expectedstudies of other sandstones will verify this result.

PREDICTING PERMEABILITY

Permeability values observed among a set of sam-ples may not include the maximum value likely tooccur, because of incomplete sampling, incompletecore recovery, or the well bore missing the maximumdevelopment of porosity in a depositional subfacies.Given the Hagen-Poiseuille permeability model(Ehrlich et al., 1991b), however, permeability can becalculated for a series of “synthetic” rocks that cancontain pore type proportions and porosities exceed-ing those in the sample set, but falling within plausi-ble limits, as discussed in the example from thePattani and Cherokee basins.

Detecting Permeability Gradients in Sandstone Complexes––Quantifying the Effect of Diagenesis on Fabric 107

Page 122: Reservoir Quality Prediction in Sand and Carbonates

108 Ehrlich et al.

Diagenetic changes are responsible for the changesin throat size associated with each pore type. Reduc-tion in throat sizes associated with the loose-packedcircuits in sandstones in the following examples can beof two types: (1) progressive development of quartzovergrowths can gradually reduce the throat size (thecircuits remain intact) or (2) local patches of diageneticcarbonate or clay can plug the circuits. Both situationscan be modeled, with the case involving intact circuitsproviding the most optimistic picture. As discussed inthe following examples, the reduction in throat size

associated with intact circuits can change smoothly asa function of depth, defining a diagenetic gradient,which in turn controls the maximum permeability thatmay be encountered.

Example 1: Satun Field, Pattani Basin, Gulf of Thailand

An extensive coring program in the Satun field, Pat-tani Basin in the Gulf of Thailand provided the oppor-tunity to sample Miocene sands over a depth range of~1000 m (Bowers et al., 1994). The sandstones share a

Figure 4. Relationship ofpermeability to depth inMiocene sandstones, Gulf ofThailand. Squares representsamples with permeabilityand porosity data only (n =197). Dots represent the sub-set of samples with mercury-injection and image analysisdata.

Figure 3. Relationshipbetween grain size and per-meability of the PerrySandstone; open circles rep-resent samples with patchycarbonate cement; squaresrepresent samples with uni-formly distributed quartzovergrowths. Slope of theregression line through theuniformly affected sampleyields an exponent of ~5compared with exponents inthe range of 1.3 to 2 reportedby Shepard (1989) (solid lineto left of data).

Page 123: Reservoir Quality Prediction in Sand and Carbonates

common depositional environment and have approxi-mately the same grain size. Maximum permeabilitydecreases with depth from >1 darcy (6000 ft) to <10 md(8500 ft) (Figure 4). The samples come from a “hot” basinwhere a high geothermal gradient (4.0–5.0°C/100 m) isassociated with a diagenetic gradient as feldspars areprogressively destroyed as a function of depth (Tra-vena and Clark, 1986). Quartz overgrowth develop-ment and kaolinitization developed in step withfeldspar dissolution. A hypothesis of the study wasthat the diagenetic gradient would be reflected in thepore type–pore throat size relationship as a function ofdepth. As discussed below, this was the case, but in anunanticipated mode.

Image analysis data were linked with permeabilityand mercury-injection data by Bowers et al. (1994),

using the procedures described earlier in this chapterand detailed in Ehrlich et al. (1991a, b) and McCreeshet al. (1991). They derived five pore types (Figure 5).Pore type 1 (PT1; mean diameter of 19 µm) is the small-est pore type and occurs as cuspate to triangularshaped porosity elements in thin section. Pore type 2(PT2; mean diameter of 37 µm) represents the survivingremnants of intergranular porosity bounded by quartzovergrowths. Pore type 3 (PT3; mean diameter of 53µm) is associated with kaolinite, which can reduce theeffective throat size. The two largest pore types aretypes 4 and 5 (PT4 and PT5) and have mean diametersof 76 µm and 160 µm, respectively. These two pore typesoccur as discrete patches of porosity surrounded by amore compact fabric and overgrown grains. Poretypes 4 and 5 are associated with enhanced permeabil-

Detecting Permeability Gradients in Sandstone Complexes––Quantifying the Effect of Diagenesis on Fabric 109

Figure 5. Pore types derivedfor the Miocene sandstones,Gulf of Thailand. Five poretypes were derived, rangingin size from 19 to 160 µm indiameter. See text for a com-plete description of each poretype. Horizontal line equals100 µm; arrows indicate anexample of each pore type.

Page 124: Reservoir Quality Prediction in Sand and Carbonates

110 Ehrlich et al.

ity (Figure 1C, D). In most samples, PT4 is the majorcarrier of permeability; PT5 was too low in abundanceto significantly affect permeability in all but a fewsamples.

Permeability modeling using the Hagen-Poiseuillemodel shows that fabric dominated by PT1, PT2, andPT3 (conventional intergranular porosity) accounts forlittle permeability throughout the depth range. Low-permeability samples anywhere in the sampled depthrange are dominated by these pore types. Therefore,the effectiveness of the conventional intergranularporosity has been impaired throughout the depthrange. Pore types 1, 2, and 3 are associated with kaolin-ite throughout the depth range, indicating that the dia-genetic gradient observed by Travena and Clark (1986)is not coincident with the gradient in the conventionalintergranular porosity. The diagenetic gradient isreflected in the distribution of PT4 and PT5, however.

Pore types 4 and 5 are unusual in that they occur inthin section as patches of large pores completely sur-rounded by PT1, PT2, and PT3. Although isolated in sec-tion view, porosity of pore types 4 and 5 must beconnected in the third dimension to account for theobserved relationships between pore type and porethroat size. PT4 and PT5 contain little, if any, kaolinite,indicating that these pores were not in existence duringthe period of kaolinite formation. Samples with PT4 andPT5 also exhibit a less compacted fabric than the fabricsdominated by PT1, PT2, and PT3. These characteristics,coupled with the patchy distribution of the large inter-granular pores associated with PT4 and PT5, led Bowers

et al. (1994) to interpret these pore types as representingthe product of a late-stage dissolution of an early patchycarbonate cement. While no vestige of this cementoccurs in the sampled rocks, carbonate cement is abun-dant in sandstones shallower in the sequence.

Throat sizes associated with PT4 decrease log-linearlyover the sampled depth range (Figure 6), accounting forthe observed reduction in maximum permeability withrespect to depth. If a pore type is assumed to exist at anydepth, the maximum permeability likely to occur at anygiven depth can be calculated. Using that assumption,the relative effectiveness of each pore type can be illus-trated by creating “synthetic” rocks, each with the sameporosity and containing a single pore type. Bowers et al.(1994) constructed such models containing throat sizesappropriate for shallow depths (Figure 7). They con-cluded that permeability values >1 darcy are possibleonly in the presence of PT4. Less than 10% of PT4 wouldensure permeability values >100 md; PT4 permeabilityefficiency decreases with depth as its throat sizedecreases. Bowers et al. (1994) showed that similar val-ues of PT4, yielding a permeability of 1 darcy at a depthof 6000 ft; would account for a permeability of ~100 mdat 7000 ft, and only about 10 md at ~8000 ft (Figure 8).These values are in agreement with the maximum mea-sured permeabilities over that depth range.

Figure 6. Relationship of mean throat radius of poretype 4 (PT4) with depth in Miocene sandstones,Gulf of Thailand. The throat radius associated withPT4 decreases log-linearly with depth.

Figure 7. Relationship between optical porosity andpermeability for rocks composed entirely of a singlepore type in Miocene sandstones, Gulf of Thailand.The relationships were calculated from the perme-ability model assuming throat sizes associated withthe most permeable zone (5560–6240 ft). Note thatonly PT4 and PT5 can account for values of perme-ability >1 md.

Page 125: Reservoir Quality Prediction in Sand and Carbonates

Example 2: Upper Carboniferous Sandstones,Cherokee Basin, Oklahoma

An industry–university consortium was organizedto evaluate measurement-while-drilling tools in verti-cal and deviated boreholes in the Cherokee Basin inNorth Central Oklahoma (Hutchinson, 1991). A by-product of this multidisciplinary investigation wasporosity, permeability, density, and other analyses ofmore than 1000 plugs from a core taken in the verticalborehole. The core spanned 2700 ft of Permian andUpper Carboniferous sedimentary rocks. Of the plugstaken by the consortium, Riggert (1994) selected 73samples, spanning >1000 ft in four Upper Carbonifer-ous (Missourian and Virgilian) sandstones. The sand-stones are medium to very fine grained, quartz-richsandstones with subsidiary amounts of feldspar andlithic fragments. Patchy carbonate cement occurs in allsamples and can be a major factor in permeabilityreduction. The samples come from a “cold” basin witha low geothermal gradient (<1.5°C/100 m).

Using image analysis, Riggert (1994) determined thatfour pore types were sufficient to account for essentiallyall of the petrographic variability in these sandstones.All pore types represent intergranular pores of varioussizes and shapes and are illustrated in Figure 9. (Note:Pore type numbers refer to the relative sizing of opticalporosity types within an individual reservoir. Thelargest pore type is identified by the largest pore type

number. A pore type with same number designation inone reservoir is not related to the same numericallylabeled pore type in another reservoir.) Pore type 1 hasa mean diameter of 14 µm and is characterized by com-pact intergranular pores. Pore type 2 has a mean diame-ter of 19 µm and is characterized by small elongatedintergranular pores. Pore types 3 and 4 are the twolargest pore types and have mean diameters of 39 µmand 79 µm, respectively; they are associated with loose-packed domains and are the primary agents forenhanced permeability. In some samples, the presenceof these pore types does not ensure enhanced perme-ability because circuits associated with these pore typesare blocked by carbonate cement. When this occurs, thepore throat radii associated with PT3 and PT4 decreasefrom 6–10 µm to <3 µm.

Using only samples with open circuits, permeabilityvalues are depth related, because the throat sizes of PT3and PT4 decrease with depth. From shallowest to deep-est, throat radii of these two pore types are ~10 µm in theHoover Sandstone, ~8 µm in the Elgin Sandstone, ~6 µmin the Perry Sandstone, and ~5 µm in the Layton Sand-stone. Assuming a maximum core porosity of 20% anda maximum relative proportion of PT3 of 50%, the max-imum permeabilities for the sandstones in this sequencecan be calculated (Figure 10A). From this relationship,the maximum permeability in the sandstone betweenthe Elgin and Perry sandstones, the Tonkawa [sampledby the consortium, but not by Riggert (1994)] can beinterpolated and compared with measurement (Figure10B).

DISCUSSION AND CONCLUSIONS

Permeability is dependent on grain size, packing, sort-ing, and diagenetic state. In the absence of diagenesis,permeability prediction becomes an exercise in predict-ing depositional fabric with respect to basin location(i.e., facies distribution and burial history analysis).Holding depositional fabric constant, permeabilityvaries in response to changes in diagenetic state.

Quantifying the diagenetic state relevant to perme-ability prediction is difficult to impossible at presentbecause diagenesis is a combination of the effects ofphysical and chemical processes. Many geochemicalchanges involve the physical redistribution of phases.However, diagenetic processes invariably affect theporosity: pores change in size and shape, pore throatsizes change, or the relationship between pore sizesand pore throat sizes changes. Therefore, these charac-teristics (which are quantifiable) can be used for adirect characterization of diagenetic state (which is notquantifiable) for other rocks in the basin.

Most of these changes in the sizes of pores and porethroats can be detected by using the procedures ofpore type determination described in Ehrlich et al.(1991b), and for relating pore types to pore throat sizedescribed by McCreesh et al. (1991). Building on this,permeability can be partitioned among pore types byusing the methods described by Ehrlich et al. (1991a);the porosity elements responsible for enhanced per-meability can be identified by this method. Examples

Detecting Permeability Gradients in Sandstone Complexes––Quantifying the Effect of Diagenesis on Fabric 111

Figure 8. Relationship between optical porosityand permeability of modeled rocks composed ofPT4 with throat radii appropriate for depths of6000 ft, 7000 ft, and 8000 ft for Miocene sandstones,Gulf of Thailand.

Page 126: Reservoir Quality Prediction in Sand and Carbonates

112 Ehrlich et al.

discussed in this chapter showed that pore throat sizesassociated with such pore types vary smoothly as afunction of depth in two basins, one with a high geothermal gradient and one with a low geothermalgradient. The maximum permeability associated withthese fabric types can be interpolated within the depthrange. With care, extrapolation may be made togreater depths and laterally away from well control.

The results described here do not require that themaximum permeability be measured in availablecore, only that the pore types associated with perme-ability enhancement be present in a few samples. Thismeans that the maximum measured permeabilitymay be unrepresentative of what may be possible in awell bore; higher permeabilities may be encounteredwith additional drilling. On the other hand, the maxi-mum likely permeability may be that which was mea-sured; if that permeability is below the economicthreshold permeability, reservoir quality may be toolow for exploitation.

Our results until now describe changes withrespect to depth. A logical next step is to attempt per-meability prediction using data taken from reservoirscovering a wider areal extent. A potential shortcut isan attempt to relate basin history models (especiallythose incorporating heat flow over time) with the dia-genetic gradient expressed by the changes in the porethroat size of the pore type that carries the majority ofthe permeability.

An unresolved aspect of permeability prediction isthe degree of reduction of average permeability as afunction of diagenesis. Decreased average permeabil-ity independent of grain size is generally the productof detrital mineral composition, burial depth, over-pressure, temperature, and cement type. In the case ofpatchy carbonate cements, they progressively isolateportions of the loose-packed porosity that are respon-sible for much of the permeability in unaffected sand-stones. The early stages of such mineralization may bebenign; Prince et al. (1995) observed that there is apreference for mineralization of close-packeddomains, accounting for the reduction of pore throatsizes compared to those in the loose-packed domains.However, given a great enough chemical potential,such mineralization can overcome the effect of fabricstructure and reduce the permeability by blocking theloose-packed circuits.

In our experience, permeability reduction is com-monly associated with the progressive nucleation andgrowth of patchy carbonate cement. We do not know atthis time whether the degree of this kind of cementa-tion has a large-scale spatial (basinal) component orwhether it is essentially controlled by local factors. Inthe case of the Oklahoma sandstones, carbonatecementation becomes more pervasive with depth.Schmidt and McDonald (1979) report a tendency forcementation in some basins to decrease with depth, so

Figure 9. Pore types derivedfor the Upper Carboniferoussandstones of Oklahoma.Four pore types werederived, ranging in size from14 to 79 µm in diameter. Seetext for a complete descrip-tion of each pore type. Eachview is 1075 ×832 µm.

Page 127: Reservoir Quality Prediction in Sand and Carbonates

Detecting Permeability Gradients in Sandstone Complexes––Quantifying the Effect of Diagenesis on Fabric 113

Figure 10. Relationship betweenpermeability and depth, UpperCarboniferous sandstones,Oklahoma. Squares representmeasured values; solid dots andthe connecting line representmaximum permeability calculat-ed according to the permeabilitymodel. (A) Samples taken byRiggert (1994); (B) samples takenby the industry–academic consortium in a similar depthinterval (n = 431).

Page 128: Reservoir Quality Prediction in Sand and Carbonates

114 Ehrlich et al.

there is some hope that such cementation can be under-stood and quantified, permitting estimates of averagepermeabilities. In addition, as shown in the examplefrom the Gulf of Thailand, relatively late stage dissolu-tion of this cement can restore permeability after theconventional fabric is rendered ineffective by diagene-sis. The possibility of such a restoration of porosity andpermeability must be kept in mind in order not toexclude from consideration reservoirs whose conven-tional intergranular porosity would be predicted not tosupport high values of permeability.

The gradients described in this chapter imply acontinuity of history from shallow to deep strata.Unconformities, faults, and other barriers may causediscontinuities in the rate of change of throat sizesand will make extrapolation more complicated, butstill possible with the techniques described.

ACKNOWLEDGMENTS

This manuscript has been significantly improved byreviews from Jon Gluyas, Andy Bradshaw, JulieKupecz, and Sal Bloch.

REFERENCES CITED

Anguy, Y., R. Ehrlich, C.M. Prince, V.L. Riggert, andD. Bernard, 1994, The sample support problem forpermeability assessment in sandstone reservoirs, inJ. M. Yarus and R.L. Chambers, eds., Stochasticmodeling and geostatistics: AAPG ComputerApplications in Geology 3, p. 37–54.

Beard, D.C., and P.K. Weyl, 1973, Influence of textureon porosity and permeability of unconsolidatedsand: AAPG Bulletin, v. 57, no. 2, p. 348–369.

Bowers, M.C., 1992, The use of nuclear magnetic res-onance, permeability and diffusion to characterizethe porous microstructure of sandstones: Ph.D.thesis, University of South Carolina, Columbia,South Carolina, 152 p.

Bowers, M.C., R. Ehrlich, and R.A. Clark, 1994, Deter-mination of petrographic factors controlling perme-ability using image analysis and core data, SatunField, Pattani Basin, Gulf of Thailand: Marine andPetroleum Geology, v. 11, no. 2, p. 148–156.

Ehrlich, R., E.L. Etris, D. Brumfield, and L.P. Yuan,1991a, Petrography and reservoir physics III: physi-cal models for permeability and formation factor:AAPG Bulletin, v. 75, no. 10, p. 1579–1592.

Ehrlich, R., K.O. Horkowitz, J.P. Horkowitz, and S.J.Crabtree, 1991b, Petrography and reservoir physicsI: objective classification of reservoir porosity:AAPG Bulletin, v. 75, no. 10, p. 1547–1562.

Evans, J.C., R. Ehrlich, D. Krantz, and W.E. Full, 1992,A comparison between polytopic vector analysisand empirical orthogonal function analysis for ana-lyzing quasigeostrophic potential vorticity: Jour.Geophys. Res., v. 97, no. C2, p. 2365–2378.

Fraser, H.J., 1935, Experimental study of the porosityand permeability of clastic sediments: Jour. Geol.,v. 43, no. 8, p. 910–975.

Full, W.E., R. Ehrlich, and J.E. Klovan, 1981, ExtendedQModel—objective definition of external endmembers in the analysis of mixtures: J. Math. Geol.,v. 13, no. 4, p. 331–344.

Graton, L.C., and H.C. Fraser, 1935, Systematic packingof spheres with particular relation to porosity andpermeability: Jour. Geol., v. 43, no. 8, p. 785–909.

Horkowitz, K.O., 1987, Direct and indirect control ofdepositional fabric on porosity, permeability, andpore size geometry: differential effect of sandstonesubfacies on fluid flow, Cut Bank Sandstone, Mon-tana: Ph.D. thesis, University of South Carolina,Columbia, South Carolina, 136 p.

Hutchinson, M.W., 1991, Comparisons of MWD, wire-line and core data from a borehole test facility, 66thAnnual Technical Conference and Exhibition: SPEPaper 22735, p. 741–754.

McCreesh, C.A., R. Ehrlich, and S.J. Crabtree, 1991, Pet-rography and reservoir physics II: relating thin sec-tion porosity to capillary pressure, the associationbetween pore types and throat size: AAPG Bulletin,v. 75, no. 10, p. 1563–1578.

Murray, C.J., R. Ehrlich, E. Mason, and R. Clark, 1994,Evaluation of the diagenetic and structural influenceson hydrocarbon entrapment in the Cardium Forma-tion, Deep Basin, western Alberta: Bulletin of Cana-dian Petroleum Geology, v. 42, no. 4, p. 529–544.

Prince, C.M., R. Ehrlich, and Y. Anguy, 1995, Analysisof spatial order in sandstones II: grain clusters,packing flaws, and the small-scale structure ofsandstones: Jour. Sed. Res., v. A65, no. 1, p. 13–28.

Riggert, V.L., 1994, Petrophysical relationships ofpores and pore throats to spatial fabric elements insandstones and their implications for fluid and elec-trical flow: Ph.D. thesis, University of South Car-olina, Columbia, South Carolina, 192 p.

Schmidt, V., and D.A. McDonald, 1979, The role of sec-ondary porosity in the course of sandstone diagene-sis, in Aspects of Diagenesis: SEPM SpecialPublication 26, p. 175–207.

Shepard, R.G., 1989, Correlations of permeability andgrain size: Groundwater, v. 27, no. 5, p. 633–638.

Travena, A.S., and R.A. Clark, 1986, Diagenesis of sand-stone reservoirs of Pattani Basin, Gulf of Thailand:AAPG Bulletin, v. 70, p. 299–308.

Page 129: Reservoir Quality Prediction in Sand and Carbonates

115

Chapter 9

Geostatistical Simulation of ReservoirPorosity Distribution from 3-D, 3-C Seismic

Reflection and Core Data in the Lower NiskuFormation at Joffre Field, Alberta

Raúl Cabrera-GarzónJohn F. ArestadKadri DagdelenThomas L. Davis

Department of Geophysics, Colorado School of MinesGolden, Colorado, U.S.A.

ABSTRACT

Rock properties such as lithology and porosity can be obtained from com-parative P- and S-wave traveltimes or velocities measured from multicom-ponent (3-D, 3-C) seismic reflection data. A 3-D, 3-C seismic reflection datasurvey was acquired by the Colorado School of Mines ReservoirCharacterization Project at Joffre field, Alberta, to map the complex porositydistribution in a shelf carbonate reservoir. Velocity ratio analysis, of com-pressional velocity to shear velocity (Vp/Vs), indicates a linear correlationwith porosity in the Devonian Nisku reservoir. Vertical porosity distributionat wells and horizontal porosity distribution derived from seismic reflectiondata are used to map 3-D porosity distribution using geostatistical methods.The results show enhanced mapping of porosity distribution and better defi-nition of the lateral limits of the reservoir. These results will assist in reser-voir simulation of this field.

INTRODUCTION

An accurate determination of the spatial distributionof porosity is key to understanding and predictingpetroleum reservoir performance. The information thatcan be used to characterize porosity distribution isdiverse. For instance, well information provides goodvertical resolution; however, it gives poor horizontalresolution due to the large separation between wells.On the other hand, seismic reflection data provide highhorizontal resolution but lower vertical resolution thandoes well information. Geostatistical tools are useful in

relating different types of rock property measure-ments, such as wireline logs, core measurements, andseismic reflection data, to provide models that describethe spatial distribution of the properties being esti-mated. Significant porosity differences occur in thelower Nisku interval at Joffre field, where a 3-D, 3-Cseismic reflection data survey was acquired to provideP- and split shear wave data [(fast) S1 and (slow) S2] .Shear wave splitting is considered to occur due to dif-ferential horizontal stress, fracturing, and pore shapeelongation. Conventional P-wave seismic reflectiondata have been ineffective for porosity characterization

Cabrera-Garzón, R., J.F. Arestad, K. Dagdelen, andT.L. Davis, 1997, Geostatistical simulation of reser-voir porosity distribution from 3-D, 3-C seismicreflection and core data in the Lower NiskuFormation at Joffre Field, Alberta, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality pre-diction in sandstones and carbonates: AAPGMemoir 69, p. 115–125.

Page 130: Reservoir Quality Prediction in Sand and Carbonates

116 Cabrera-Garzón et al.

of this reservoir. Geostatistics is used to analyze thediverse geophysical data, to develop useful relation-ships among the different types of information, and tocharacterize porosity distribution.

POROSITY PREDICTION FROMSEISMIC REFLECTION DATA

The motivation for using seismic reflection data tocharacterize the spatial distribution of porosity (orother physical properties) comes from the ability toprovide useful relationships between the seismicreflection data and physical properties. As will beshown later, porosity–seismic data relationships havebeen developed both theoretically and experimentally,mainly for sand and sand-clay models, but little workhas been done on carbonate reservoirs.

Traditionally, velocity–porosity relationships havebeen estimated from regression methods that fit lineartrends for certain intervals. More recently, the use of sta-tistical techniques has provided better results in relatingporosity to seismic reflection attributes, particularly fordescribing interwell porosity from surface seismic reflec-tion data. Doyen (1988) applies geostatistical techniquesto relate transit times from surface seismic reflection toporosity measurements from wells, and compares theresults to those derived from linear regression. Scerboand Mazzotti (1991) apply cokriging methods to relateseismic velocities to porosity. This approach providesbetter results than those provided by kriging methods.However, the results still show the smoothing and holeeffects imposed by the original kriging method.

More recent approaches use simulation techniquesthat provide porosity models that describe the spatialdistribution of this property. Such models are stronglysupported by both statistical models that correlateporosity and seismic reflection attributes and by theinformation itself (Deutsch and Journel, 1992).

The first approaches to estimate porosity from seis-mic reflection data have considered changes in com-pressional and shear velocity due to this property.Experimental relationships among velocity, porosity,and clay content have been described by Wyllie et al.(1956, 1962), Eberhart-Phillips et al. (1989), Klimentos(1991), Marion et al. (1992), and Mavko and Nolen-Hoeksema (1992).

Davis et al. (1992) provide results that were derivedfrom detailed three-dimensional, multicomponent(3-D, 3-C) seismic reflection data. Such results showthe potential to relate anisotropy and porosity toVp/Vs ratios and shear velocity differences.

Vernik and Nur (1992) presented work relatingpetrophysics to porosity and velocity. They developeda petrophysical classification of siliciclastics to predictlithology and porosity from seismic velocities. Theypresented results for Vp/Vs vs. porosity. In their workthey fit linear and polynomial trends to the laboratorydata; their results show an increase of Vp/Vs withporosity. The fits for arenite and clean arenite are poly-nomial, whereas the fits for the case of shale andwackestones are linear (Figure 1).

Berge et al. (1995) established an excellent agree-ment between compressional and shear velocities fromlaboratory and theoretical predictions of velocity frombounded methods. Figure 2 shows the Vp/Vs–porosityrelationship estimated from their numerical results,which is good for the case of wet samples. On the otherhand, a Vp/Vs–porosity relation cannot be estimatedfor the dry samples. As in Vernik and Nur’s (1992)Vp/Vs–porosity relationship, these Vp/Vs results (forthe wet case) also increase with increasing porosity.

Sarmiento (1994) discussed the usefulness of theVp/Vs ratios as a tool for identifying lithology, andalso established that Vp/Vs vs. Vp plots of the Niskureservoir are not constant relationships but vary withporosity. Therefore, Vp/Vs ratios can be used not onlyfor lithology discrimination, but also for porosity map-ping. From wireline log data, Sarmiento proposed thevalues 2.0 for anhydrite and 1.87 for dolomite for theNisku Formation. One of the biggest concerns whenanalyzing the Vp/Vs1 data was that high Vp/Vsratios were expected in the NE part of the seismic areaand lower Vp/Vs values in the SW, according to theknowledge of anhydrite distribution in the field. How-ever, the trend from the Vp/Vs1 ratio map shows theopposite relationship.

If we consider that porosity increases with Vp/Vsratio in the dolomite zone, and that porosity is almostzero due to anhydrite plugging, then the increasingtrend that we observe in our data is valid if a superpo-sition of effects is considered. Figure 3 illustrates theidea that validates the use of Vp/Vs not only for lithol-ogy discrimination but also for porosity estimation.

Also to be considered is that the values given bySarmiento (1994) are for pure anhydrite and puredolomite. In the Nisku reservoir, anhydrite has

0 5 10 15 20 25 30 35Porosity (% )

1.5

1.7

1.9

2.1

Vp

/Vs

W acke

ShaleArenite

Clean Arenite

Pe = 40 MPa

Figure 1. Vp/Vs ratios for saturated rocks vs. porosity(after Vernik and Nur, 1992). Pe = effective pressure,MPa = megapascals.

Page 131: Reservoir Quality Prediction in Sand and Carbonates

3-D, 3-C Seismic Reflection and Core Data, Lower Nisku Formation 117

0.0 0.1 0.2 0.3 0.4 0.5Fractional Porosity

1.50

1.75

2.00

Vp

/Vs

0.0 0.1 0.2 0.3 0.4 0.5Fractional Porosity

1.60

2.00

Vp

/Vs

Vp/Vs vs. Porosity (w et sam ples)

V p / V s v s . P o r o s i t y ( d r y s a m p l e s )Figure 2. Vp/Vs vs. porosityfor sandstone analogs. Dryand wet cases (data fromBerge et al., 1995).

D olom ite whe re a nhydritehas p lug ged porosity(N on porous zone)

D olom ite(P orous zone)

V p /V s response due to litho logy

V p /V s response due to po rosity changes

V p /V s response due to litho logy + po rosity changes

P oros ity increases Zone o fze ro porosity

Vp

/Vs

incr

ea

ses

Figure 3. Qualitative interpreta-tion of Vp/Vs ratio changes seenas a combination of responsesdue to lithology and porositychanges.

Page 132: Reservoir Quality Prediction in Sand and Carbonates

118 Cabrera-Garzón et al.

plugged porosity in dolomite rock; thus, the non-porous zone will consist of a mixture of dolomiteand anhydrite. If this is true for the reservoir, the dif-ference of Vp/Vs ratios between dolomite anddolomite+anhydrite will be very small, and Vp/Vschanges due to porosity will predominate. Anotherimportant point to justify the use of Vp/Vs1 ratiosfor porosity estimation is that reservoir productionand pressure data support the conclusion thatVp/Vs ratio variations over the Nisku reservoir aredue to porosity, and not changing reservoir fluids(Arestad, 1995).

Reservoir pressures within the limits of the 3-D, 3-Csurvey were above the bubble point pressure at thetime of data acquisition, showing that no free reservoirgas is located in the survey area (Al-Bastaki et al., 1995).

JOFFRE FIELD

Joffre field is located on the inner Nisku shelf regionof South-Central Alberta, between Calgary andEdmonton, at the western edge of the Bashaw Complex(Figure 4). The field covers townships 39 and 38 andranges 27 and 26, for an areal extent of 45–50 mi2

(116.5–130.5 km2). The geological model of the Niskuinterval has been continuously changed as knowledgeof the field has increased. This area is located in thecentral part of the Phanerozoic Western Canada Sedi-mentary Basin. The Devonian sedimentary section con-sists of marine carbonates, shales, and evaporites(Al-Bastaki et al., 1995). The Upper Devonian in theWestern Canada Sedimentary Basin has been dividedinto four groups, from oldest to youngest: the Beaver-hill Lake, Woodbend, Winterburn, and Wabamungroups (Figure 5). The Nisku Formation of the Winter-burn Group consists of two units in the Joffre area: anupper unit of dolomite interbedded with anhydrite,and a lower, open marine dolomite with vuggy poros-ity and minor anhydrite. Arestad (1995) establishedthat the stratigraphic zonation found in the openmarine unit is present, with local variations, in most ofthe cored wells. The reservoir portion of the lowerNisku beneath the seismic reflection data area has analmost constant thickness of 22 m.

GEOPHYSICAL DATA

The well information used for this study describ-ing the Nisku interval is restricted to an area of 6.5 ×5.5 km, slightly larger than the 3-D seismic data set.The area contains a total of 44 wells, including coresamples, porosity–permeability measurements, andwireline logs (Figure 6).

A total of twenty-three wells were selected for petro-physical measurements. All of these wells are locatedinside the well control area for this study. The core

Figure 4. Location of the Joffrefield study area (after Sarmiento,1994).

Figure 5. Generalized stratigraphy of the UpperDevonian in South-Central Alberta (after Watts,1987).

Page 133: Reservoir Quality Prediction in Sand and Carbonates

samples were 3.5–4 in. (7.62–10.16 cm) in diameter and5–11 in. (12.5–28 cm) long. To measure porosity andpermeability, complete cores were used rather thansmall plugs, due to the moldic and vuggy nature of theporous Nisku reservoir rock.

Figure 7 shows porosities and permeabilities vs.depth for well 09-21-39-26 to show how these proper-ties are distributed through the reservoir interval.

The seismic reflection data used in this study are partof the result of a 3-D, 3-C survey acquired over thenortheastern edge of the Nisku reservoir by the Col-orado School of Mines Reservoir CharacterizationProject. Conventional seismic reflection data have failedto characterize the complex diagenetic dolomite reser-voir. In general, the acoustic impedance contrastbetween the Nisku reservoir and surrounding rock isvery small. Additionally, strong interbed multiples andconverted waves (P-SV) can interfere with the Niskuevent on stacked data (Davis, 1992); thus, variations inthe Nisku reflection event are usually not reliable indi-cators of porosity development or of reservoir quality.Therefore, 3-D, 3-C seismic reflection data technology

has been applied to the field to improve reseroir char-acterization. The use of this technology allows therecording of compressional as well as shear wave data.Anisotropic media (like carbonate reservoirs, whichpresent azimuthal fracturing, and elongated poreshape porosity) create splitting and polarization of theshear waves (Martin and Davis, 1987). The results are afast (S1) and a slow (S2) shear wave data set.

The compressional S1 and S2 shear data sets wereused by Arestad (1995) to generate velocity ratiomaps, amplitude maps, and time structure maps atseveral intervals or times. The seismic data studiedin this work consist of a map of velocity ratios(Vp/Vs1) for the D1 to mid-Ireton interval (Figure8). The map covers an area of ~4 km ×3 km, with abin size of 30 m ×30 m. The velocity ratio map wascomputed utilizing interval traveltimes from boththe compressional and shear wave data sets, usingVSP data to tie the P- and S-wave reflections origi-nated from equal depths. Arestad (1995) gives adetailed description of the timing analysis and cal-culations for Vp/Vs mapping.

3-D, 3-C Seismic Reflection and Core Data, Lower Nisku Formation 119

T39N-R26W4

3-D Survey W ell Control

313000 315000 317000 319000 321000

Easting

5799000

5801000

5803000

5805000

5807000

5809000

No

rth

ing

Area

Figure 6. Location mapshowing the limits of the3-D survey and the wellcontrol area.

Page 134: Reservoir Quality Prediction in Sand and Carbonates

120 Cabrera-Garzón et al.

GEOSTATISTICAL ANALYSISThe 3-D estimation of porosity using a sequential

Gaussian simulation requires the knowledge of boththe probability distribution and the 3-D covariance orvariogram model for the physical property (Deutschand Journel, 1992). Other techniques such as cokrigingrequire variogram models for primary and secondarydata (porosity and seismic) and the cross-variogrambetween variables. For this work, neither the vario-gram model nor the cross-variogram could be deter-mined from core porosity and seismic data. Therefore,the cokriging technique could not be used to estimateporosity distribution.

The histogram of the core porosity is shown in Fig-ure 9. The histogram is not normally distributed. Themedian is the parameter that better indicates the highof the population. Porosity values range from ~0% to20%, and most of the values are concentrated in theinterval 0%–6%. The next step in calculating statisticalparameters is to estimate the 3-D variogram modelfrom core porosity. A variogram model is representedby a function that varies with increasing distance. Themaximum of the curve is called the sill, and its magni-tude represents the variance of the data. The range isthe distance at which the sill is reached; it representsthe correlation length of the parameter being

Figure 7. Porosity and perme-ability distribution for well09-21-39-26.

-1220-1200-1180Depth

0.0

0.1

0.2

Fra

ctio

nal

Po

rosi

ty

-1220-1200-1180

0

400

800

Ver

tica

l per

mea

bili

ty(m

ilid

arcy

)

-1220-1200-1180

0

200

400

Max

imu

m h

ori

zon

tal

per

mea

bili

ty(m

ilid

arcy

)

-1220-1200-1180

0

200

400

Per

pen

dic

ula

r to

max

imu

mh

ori

zon

tal P

erm

eab

ility

(mili

dar

cy)

W ell 09-21-39-26

Page 135: Reservoir Quality Prediction in Sand and Carbonates

measured. Along the vertical direction, the sampleinterval is small enough to obtain a high-resolutionvariogram, which allows modeling of the Z-compo-nent of the variogram model. The result of this compu-tation is shown in Figure 10. On the other hand,computation of horizontal variograms for the N-S andE-W directions shows a lack of information due to thelarge spacing (~0.5 mi; 0.805 km) distance between

wells. The computed variograms cannot be modeledbecause the sill has been reached before the first lagdistance. Two variogram models are plotted along thecomputed variogram to show how the range (or dis-tance of maximum continuity) can be represented forany model (Figure 11).

It is clear that we need to estimate the horizontalvariogram model from other data. Thus, by establish-ing the relationship between core porosity and seismicreflection information (in the form of Vp/Vs or ampli-tude of shear wave maps), we can transform seismicreflection data into horizontally distributed porositydata. Considering that we have a high-resolutionporosity description in the vertical direction that is tobe compared with high-resolution horizontal seismicreflection data, we need to obtain an average (or mean)porosity at each well location. Once the mean porosi-ties were calculated, the seismic reflection values sur-rounding each well were extracted from the data set tocalculate a mean seismic reflection data attributevalue. Data within a 90-, 150-, and 250-m radiusaround each well were considered. Then, the mean

3-D, 3-C Seismic Reflection and Core Data, Lower Nisku Formation 121

Figure 8. Velocity ratio (Vp/Vs1) map of theWabamun (D1) to mid-Ireton interval.

Figure 10. Vertical core porosity variogram; the vari-ogram model is exponential.

Figure 9. Histogram of core porosity measured at23 well locations within the lower Nisku.

317000 319000 321000

5807500

5805500

Easting

Northing

1.50 1.75 2.00 2.25 2.50

0.00 0.05 0.10 0.15 0.20Fractional porosity

0

50

100

Co

un

ts

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15Distance (m)

0.0000

0.0002

0.0004

0.0006

0.0008

0.0010

0 500 1000 1500 2000 2500D istance (m )

0.0000

0.0005

0.0010

N-S direction E-W direction

Figure 11. Horizontal core porosity variograms. Notethe lack of definition due to well separation.

Fra

ctio

nal

Po

rosi

tyF

ract

ion

al P

oro

sity

Page 136: Reservoir Quality Prediction in Sand and Carbonates

122 Cabrera-Garzón et al.

Vp/Vs1 ratios were compared to the mean porositiesin the form of crossplots. Plots for radii of 90 and 250m are shown in Figure 12.

The Vp/Vs1 vs mean porosity crossplot indicatesan increase of porosity with increasing Vp/Vs1. Linearand exponential models were fitted to the data toderive a mathematical relationship between Vp/Vs1and mean porosity. There are no significant differ-ences between these two models over the examinedinterval; therefore, the linear model is chosen to trans-form Vp/Vs1 ratios into mean porosity values. Figure13 shows a map of the estimated porosities fromVp/Vs1. If Vp/Vs1 also indicates lithology, limits ofthe reservoir due to anhydrite plugging can also beinterpreted from this map.

Variograms were calculated from the porosity mapderived from Vp/Vs1 for eight directions. The resultsshow that horizontal variation of porosity is geometri-cally isotropic within a distance of 600 m. Two direc-tions (N68W and N22E) are shown in Figure 14.

Assuming that the spatial variability structures ofthe porosity from seismic reflection and core data arethe same, the variogram analysis of the seismic reflec-tion data-derived mean porosity map defines the near-lag portion of the curve that is not defined in thehorizontal core porosity variogram. The sills of bothhorizontal variograms (core porosity and porosityfrom seismic reflection data) can be related by the fol-lowing scaling relationship:

(1)

where σc = core porosity standard deviation, mc = coreporosity mean, σs = porosity from seismic reflection data

standard deviation, ms = porosity from seismic reflectiondata mean; with the computed scaling factors being

(2)

and

(3)

The combination of previous results from core datawith the results obtained from seismic reflection datagives the 3-D variogram model, which is defined by thefollowing parameters: lateral range N-S/E-W = 600 m,vertical range (depth) Z = 3–4 m, dip direction = 0.5° W,horizontal geometrical anisotropy ratio = 1.0, and verti-cal geometrical anisotropy ratio = 0.005–0.0066.

Once the 3-D variogram model was defined, theporosity distribution was computed for the 2-D and 3-D

σ c

cm

2

2 20 000190 026

0 281= =..

.

σ s

sm

2

2 20 000680 0485

0 289= =..

.

σ σs

s

c

cm m

2

2

2

2=

0.00 0.02 0.04M ean fractional porosity

1.5

2.0

2.5

Vp

/Vs1

D ata w ith in a 90- m rad ius aro und each w e ll

122

322

327

921

922

928

1122

1226

D ata w ith in a 250- m rad ius a round each w e ll

Figure 12. Vp/Vs1 vs. mean porosity. Linear andexponential fits are plotted.

317000 319000 321000

5807500

5806500

5805500

5804500

Easting

Northing

0.00 0.05 0.10

Figure 13. Mean porosity map derived from Vp/Vs1ratios (linear fit). The scale shows fractional meanporosity.

0 400 800 1200 1600Distance (m)

0.00000

0.00005

0.00010

0.00015

0.00020

0.00025

N68W N22E

Figure 14. Horizontal variogram model and comput-ed variograms from porosity from seismic reflectiondata (porosity from Vp/Vs1 ratios).

Fra

ctio

nal

Po

rosi

ty

Page 137: Reservoir Quality Prediction in Sand and Carbonates

cases. Figure 15 shows the porosity distribution at thethree well locations used for the 2-D case. Figure 16shows the result of the conditional simulation.

Finally, the sequential Gaussian technique is used toestimate a 3-D porosity distribution. The horizontalgrid defined was 100 m, whereas the vertical samplingwas 0.5 m. According to the lateral extension definedby the well control, the number of cells calculated foreach simulation is approximately 300,000. A change inthe grid size involves a significant change in the num-ber of cells of the model and, consequently, a signifi-cant change in the number of computations needed toperform the simulation.

Figure 17 shows the result of a 3-D realization. Asexpected, the simulation of the volume of porosity distri-bution shows connectivity zones, trends of porosity hori-zons, and distribution of low- and high-porosity zones.Some of the characteristics of the 3-D porosity simulationare the enhancement of the connectivity of high- andlow-porosity zones, the preservation of the dip angle ofthe porous lower Nisku interval, and a clear definition ofthe lateral limits of the reservoir zone.

An oblique slice that runs parallel to the lowerNisku interval is shown in Figure 18. This slice showshow high-porosity values are concentrated in thesouthwest portion of the model, as indicated by thewell and seismic reflection data.

CONCLUSIONS

Geostatistical techniques have been useful toderive porosity distribution from a limited amountof core data when integrated with seismic reflectiondata (in the form of Vp/Vs1 ratios). The improve-ment on the porosity model can be evaluated bymapping the model into a petroleum reservoirsimu-lation grid. The results of history matching andreservoir performance prediction will indicate quan-titatively the amount of information gained by usingestimation techniques.

Oil reservoirs with geological settings similar tothose of Joffre field exist around the world; therefore,the use of geostatistical techniques and 3-D, 3-C seis-mic reflection data are of importance to reservoircharacterization in general, and to the implicit goal ofimproved hydrocarbon recovery.

Geostatistics can be applied to the study of spatialand temporal relationships among porosity, perme-ability, and fluid saturation related to observedchanges in seismic reflection data attributes. In thiscase, reproduction of the spatial dependence of severalvariables is critical. Therefore, conditional simulationalgorithms can be generalized to join simulation ofseveral variables.

3-D, 3-C Seismic Reflection and Core Data, Lower Nisku Formation 123

Figure 15. Distribution ofcore porosity for the wellsused in 2-D porosity simulation.

-1200-1180-1160-1140D epth (m bsl)

0.00

0.05

0.10

0.15

09-2

1F

ract

ion

al p

oro

sity

-1200-1180-1160-1140

0.00

0.05

0.10

0.15

03-2

7

-1200-1180-1160-1140

0.00

0.05

0.10

0.1512

-26

-1220

-1220

-1220

Page 138: Reservoir Quality Prediction in Sand and Carbonates

124 Cabrera-Garzón et al.

ACKNOWLEDGMENTS

We want to express appreciation for direction andfinancial support for this research to the RCP (Reser-voir Characterization Project)—Phase V Industrysponsors and to the Mexican Petroleum Institute. Wealso thank an anonymous reviewer, Robert Kendall,Julie Kupecz, and the RCP research team members fortheir comments and suggestions on this work.

REFERENCES CITED

Al-Bastaki, A.R., J.F. Arestad, K. Bard, R. Cabrera-Garzón, B. Mattocks, and M.R. Rolla, 1995, Multidis-ciplinary multicomponent reservoir characterization,Joffre field, South-Central Alberta, Canada: ColoradoSchool of Mines Reservoir Characterization Project—Phase V Sponsor Meeting Notes, April 6, 1995, 256 p.

Arestad, J.F., 1995, An integrated multicomponentthree-dimensional seismic characterization of Joffrefield, Alberta, Canada: Ph.D. thesis, ColoradoSchool of Mines, Golden, Colorado, 293 p.

Berge, P.A., B.P. Bonner, and J.G. Berryman, 1995, Ultra-sonic velocity–porosity relationships for sandstoneanalogs made from fused glass beads: Geophysics, v. 60, no. 1, p. 108–119.

Davis, T.L., 1992, Reservoir Characterization Project—Phase V: CSM Proposal 3788, 24 p.

Deutsch, C.V., and A.G. Journel, 1992, Geostatisticalsoftware library and user’s guide: New York,Oxford University Press, 340 p.

Doyen, P.M., 1988, Porosity from seismic data, a geo-statistical approach: Geophysics, v. 53, no. 10, p. 1263–1275.

Eberhart-Phillips, D., D.-H. Han, and M.D. Zoback,1989, Empirical relationships among seismic velocity,effective pressure, and clay content in sandstones:Geophysics, v. 54, no. 1, p. 82–89.

Klimentos, T., 1991, The effects of porosity-permeability-clay content on the velocity of compressional waves:Geophysics, v. 56, no. 12, p. 1930–1939.

Marion, D., A. Nur, H. Yin, and D. Han, 1992, Compres-sional velocity and porosity in sand-clay mixtures:Geophysics, v. 57, no. 4, p. 554–563.

0 500 1000 1500 2000 2500

-1180-1190-1200-1210-1220

N45E

Depth (mbsl)

0.00 0.05 0.10 0.15

Fractional porosity

Well 09-21 Well 3-27 Well 12-26

Figure 18. Oblique slice from the 3-D porosity simu-lation (from the lower Nisku interval). Lighter colorsrepresent higher porosity values.

Figure 16. SequentialGaussian conditional simulation of porosity forprofile SW 45 NE.

Figure 17. Three-dimensional conditional simulationof porosity (from the lower Nisku interval) using asequential Gaussian technique. Lighter colors represent higher porosity values.

Page 139: Reservoir Quality Prediction in Sand and Carbonates

Martin, M.A., and T.L. Davis, 1987, Shear wave bire-frigence: a new tool for evaluating fractured reser-voirs: The Leading Edge, v. 6, no. 10, p. 22–28.

Mavko, G., and R. Nolen-Hoeksema, 1994, Estimationof seismic velocities at ultrasonic frequencies inpartially saturated rocks: Geophysics, v. 59, no. 2,p. 252–258.

Sarmiento, V., 1994, Petrophysical relationships fromwireline logs for seismic calibration of the DevonianNisku and Wabamun formations, Joffre field,Alberta: Master’s thesis, Colorado School of Mines,Golden, Colorado, 97 p.

Scerbo, F., and A. Mazzotti, 1991, Geostatistical esti-mates of porosity from seismic data: Bollettino diGeofisica Teorica ed Applicata, v. 33, no. 130–131,p. 85–110.

Vernik, L., and A. Nur, 1992, Petrophysical classificationof siliciclastics for lithology and porosity predictionfrom seismic velocities: AAPG Bulletin, v. 76, no. 9, p. 1295–1309.

Watts, N.R., 1987, Carbonate sedimentology anddepositional history of the Nisku Formation(within the Western Canadian Sedimentary Basin)in South Central Alberta: GSPG Second Interna-tional Symposium on the Devonian System,p. 87–152.

Wyllie, M.R.J., A.R. Gregory, and L.W. Gardner,1956, Elastic wave velocities in heterogeneous andporous media: Geophysics, v. 21, no. 1, p. 41–70.

Wyllie, M.R.J., L.W. Gardner, and A.R. Gregory, 1962,Studies of elastic wave attenuation in porousmedia: Geophysics, v. 27, no. 5, p. 569–580.

3-D, 3-C Seismic Reflection and Core Data, Lower Nisku Formation 125

Page 140: Reservoir Quality Prediction in Sand and Carbonates

127

Chapter 10

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the

Middle Jurassic Vajont Limestone, VenetianAlps, Italy: Analogs for Hydrocarbon

Reservoirs Created Through Fault-RelatedBurial Dolomitization

William G. Zempolich1

Lawrence A. HardieDepartment of Earth and Planetary Sciences, The Johns Hopkins University

Baltimore, Maryland, U.S.A.

ABSTRACT

The Middle Jurassic Vajont Limestone of the Venetian Alps, Italy, is pre-dominantly composed of resedimented ooids that were deposited in slopeand basin settings. The Vajont Limestone has been partly replaced by mas-sive dolomite that can be mapped at both regional and local scales. Dolomitebodies that are present within or are associated with the Vajont Limestoneinclude: (1) a large-scale wedge, ~25 km long, 10–15 km wide, and ≥400–500m thick (50–94 km3), located on the hanging wall of the Alpine-aged, thrust-based Mt. Grappa–Visentin anticline. This dolomite body is located withinthe axis of the anticline and crosscuts the stratigraphic section where subver-tical to vertical faults penetrate the crest of the anticline; (2) Isolated, rootlessplume-shaped bodies, 100–200 m wide and >300 m high (≥2 ×10–2 km3),which penetrate a footwall syncline within an Alpine-aged thrust sheet.These dolomite “plumes” possess extensively brecciated cores and exhibitsharp to gradational transitions with surrounding Lower to Middle Jurassicbasinal limestone; (3) Isolated dolomite “towers” that have partly replacedCretaceous-age synsedimentary fault breccia. These bodies are found inoverlying basinal strata (i.e., the Fonzaso Formation, the Ammonitico Rosso,and the Biancone Formation), but emanate from the underlying dolomitizedVajont; and (4) Small-scale wedge-shaped dolomite bodies on the scale ofmeters found along small faults and fractures. The connection between thesedolomite bodies and Alpine-aged faults and fractures clearly indicates thatdolomitization was a late burial process.

Zempolich, W.G., and L.A. Hardie, 1997, Geometryof dolomite bodies within deep-water resedimentedoolite of the Middle Jurassic Vajont Limestone,Venetian Alps, Italy: analogs for hydrocarbon reser-voirs created through fault-related burialdolomitization, in J.A. Kupecz, J. Gluyas, and S.Bloch, eds., Reservoir quality prediction insandstones and carbonates: AAPG Memoir 69, p. 127–162.

1Present address: Mobil Oil Company, Dallas, Texas, U.S.A.

Page 141: Reservoir Quality Prediction in Sand and Carbonates

128 Zempolich and Hardie

INTRODUCTION

Dolomites constitute some of the best-quality reser-voirs for oil and gas, due to several unique properties,much as intercrystalline pore space resulting in highpermeability and resistance to burial compaction.Therefore, the prediction of dolomite body geometriesis of paramount importance in reservoir exploitation.A key to understanding dolomite distribution lies inunderstanding its origin and timing.

The origin of massive replacement dolomite hasremained one of the major unresolved problems of sed-imentology and sedimentary geochemistry for morethan a century (van Tuyl, 1916; Morrow, 1982a, b; Land,1985; Machel and Mountjoy, 1986; Hardie, 1987). In thelast two to three decades the favored interpretations forthe origin of massive dolomites have centered on“early” low-temperature replacement of limestones andlime sediments. Early low-temperature models involvesurface or near-surface marine waters such as refluxingsabkha brines or coastal mixing zone brackish waters

and thus circumvent the magnesium “supply” problem(Morrow, 1982a, b). Hardie (1987) has pointed out someof the serious weaknesses and uncertainties in theselow-temperature models for dolomitization, and hasargued that attention be turned to the many alternativeways by which massive dolomites can be made. Alongthese lines, it is valuable to compile well-documented,clear-cut case histories of as many different modes ofdolomitization as can be identified.

In this regard, the occurrence of massive replace-ment dolomite in the Vajont limestone is particularlynotable because of the deep-water slope and basinsetting in which the Vajont sediments were deposited(Bosellini et al., 1981). Features that make the Vajontarea especially valuable for field and laboratorystudy of reservoir development and predictioninclude:

1. The Vajont area is massively, but not completely,dolomitized, so that dolomitization fronts (Wilsonet al., 1990) can be mapped and provide directclues to the subsurface pathways that dolomitizing

It is proposed that during the Alpine deformation event, convection-drivenfluids derived from Late Tertiary seawater were circulated through subaque-ous Alpine-aged faults and fractures and paleosynsedimentary breccias, thuscreating the multitude of dolomite bodies now found in the Vajont and otherMesozoic basinal sediments. Paleogeographic, tectonic, and hydrologic sys-tems, similar to the one proposed for dolomitization of the Vajont, appear tobe active in modern subaqueous thrust zones of the Caribbean andNorthwest Pacific Coast.

Potential reservoir attributes of Vajont dolomite bodies include their largesize and medium to coarsely crystalline replacement fabric that is character-ized by significant amounts of partial moldic, intercrystalline, and vug porespace. Visual estimates of porosity within dolomitized grainstone and pack-stone range up to 10% to 15%, with inferred permeabilities of 1–100 md.Permeability of Vajont dolomite replacement fabrics is enhanced throughrecrystallization and the formation of touching-vug networks (inferred per-meabilities ≥100 md).

Results of this study indicate that (1) massive replacement dolomitizationin thermotectonic (i.e., burial) settings may be much more important thanpreviously thought, and (2) significant reservoirs may be hosted in other-wise tight basinal limestones as the result of late-stage burial dolomitiza-tion. Consequently, the geometries of the Vajont dolomite bodies mayprovide analogs for reservoir characterization and new exploration plays inthe subsurface. Exploration methods for analogous dolomite reservoirs in thesubsurface may include the mapping of dolomitization fronts using core andlog data and seismic reflection identification of crosscutting dolomite bodies.The focus of such efforts should be placed on anticlinal and synclinal struc-tures within buried fold and thrust belts, and along zones of deep-seated tec-tonic fractures and faults within intracratonic basins.

Page 142: Reservoir Quality Prediction in Sand and Carbonates

fluids may have taken. Outcrops are abundant andwell exposed, allowing mapping at scales thatrange from centimeters (reservoir scale) to kilome-ters (exploration scale).

2. The part of the Vajont Limestone that hosts thedolomite bodies covers an area of >100 km2, sothat a regional-scale fluid flow system capable ofproducing 50–100 km3 of dolomite must havebeen involved.

3. The Vajont consists of bedded grainstones andpackstones composed of shallow-water ooids(and some skeletal remains of shallow-waterorganisms) that were resedimented in deepwater by slope processes (Bosellini et al., 1981;Zempolich, 1995). This unusual oolite is an inte-gral part of a thick, conformable Jurassic succes-sion of deep-water sediments that were neverexposed to shallow-water syndepositional diage-netic processes. Thus, shallow-water and land-surface-related dolomitization processes can beruled out.

4. The ooids, skeletal grains, and their intergranularcements are beautifully preserved in theundolomitized parts of the Vajont area so that itis possible to document in detail their predolomi-tization petrography and isotope geochemistry;thus, the changes produced by dolomitizationcan be identified and measured.

5. Exposures of partly and massively dolomitizedVajont sediments in road cuts reveal clearly thatdolomitization fluid pathways were controlledby fractures and faults of Late Tertiary age(Alpine orogeny), indicating that the dolomitiza-tion was of “late” burial (synfaulting to postfault-ing) origin. Through an understanding of thetiming of Vajont dolomitization, the potentialexists for the “prediction” of reservoirs in othersimilarly deformed carbonate strata.

Although a number of workers have put forward evi-dence and arguments for “late” elevated-temperaturedolomitization during burial [Jodry, 1969; Zenger, 1976,1983; Mattes and Mountjoy, 1980; Broomhall and Allen,1985; Gregg, 1985; Barrett, 1987; Lee and Friedman, 1987(and the discussion of this paper by Kupecz et al., 1988,and the reply by Lee and Friedman, 1988); Aulstead etal., 1988; Zenger and Dunham, 1988; Machel and Ander-son, 1989; Cervato, 1990; Wilson et al., 1990; Kupecz andLand, 1991; Mountjoy and Halim-Dihardja, 1991; Zem-polich and Hardie, 1991a, b; Amthor et al., 1993; Dix,1993; Coniglio et al., 1994; Miller and Folk, 1994; Mon-tañez, 1994; Mountjoy and Amthor, 1994; Yao andDemicco, 1995; Zempolich, 1995], burial dolomitizationremains a controversial process believed by many sedi-mentologists to be of little importance in the origin ofancient massive dolomites (Blatt, 1982; Morrow, 1982b;Wilkinson and Algeo, 1989) and limited to the enhance-ment of preexisting or poor reservoirs (Sun, 1995). How-ever, most, if not all, of the kinetic problems of dolomiteformation that plague low-temperature systems essen-tially disappear at the elevated temperatures of burial(Hardie, 1987), making massive burial dolomitization a

likely, if not common, process. Through a comparativestudy of the Vajont limestone and dolomite, a muchclearer understanding of burial dolomitizationprocesses has been developed, including the role ofdeeply circulated subsurface fluids, in the origin ofregional-scale and isolated dolomite bodies and the evo-lution of porosity through dolomitization. These resultslead to prediction of dolomite reservoir geometries thatcan be created through tectonic and burial diageneticprocesses.

THE VAJONT LIMESTONE AND ITSGEOLOGIC SETTING

Deposition of the Vajont Limestone is closely asso-ciated with the breakup of Pangea, during which timeprolific oolite was deposited along the margins of theTethys Ocean in the circum-Mediterranean region(Bosellini, 1989; Zempolich, 1995). In the Early Juras-sic, Europe and northern Africa began to separate,and by the Late Jurassic an extensive transform zonewas present (Weissert and Bernoulli, 1985). Thebreakup of Pangea established a horst-and-grabentectonic setting along the southern Tethyan margin,and led to the structural definition of local platformsand basins (Figure 1). The Trento Platform, the mostlandward horst block of the Southern Alps, wasbounded to the west by the Lombardy Basin, and tothe east by the Belluno Basin, which separated theTrento Platform from the Friuli Platform (i.e., the sta-ble foreland). Thick sequences of these Mesozoic plat-form and basin carbonates, now partly to completelydolomitized, are extensively exposed in the VenetianAlps (Figures 2, 3).

Age dating of the Vajont Limestone is problematic.Stratigraphic-age constraints from formations locatedbelow and above the Vajont area suggest a general agerange of Bajocian to Callovian (Casati and Tomai,1969; Bosellini et al., 1981). The biostratigraphic studyof Casati and Tomai (1969) suggests an age assignment(in part) of Upper Bajocian–Lower Bathonian for theVajont limestone, based on overlapping ranges of theforaminiferal zones Protopeneropolis striata and Tro-cholina. New age constraints provided by nannofossiland ammonite data collected during the present studysuggest that the Vajont Limestone was deposited dur-ing the latest Aalenian to the earliest Bajocian (Zem-polich, 1993, 1995).

The Vajont Limestone is a particularly interesting car-bonate deposit because it is a thick sequence (≤600 malong the platform margins) predominantly composedof shallow-water oolitic sand and biogenic skeletaldebris that was redeposited by gravity flow processes inslope and basin environments (Bosellini et al., 1981;Zempolich, 1995). Depositional units include meter-scale debris flows and turbidites and beddedhemipelagic mudstone. Paleogeographic reconstruc-tions suggest that the Vajont Limestone is an eastward-thickening wedge with a depositional area in excess of100 km along strike and 50 km across strike (Figure 4).Vajont ooids were derived from the western edge of the

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 129

Page 143: Reservoir Quality Prediction in Sand and Carbonates

130 Zempolich and Hardie

Friuli Platform and were deposited as a carbonate slopeapron in the Belluno Trough (Zempolich, 1995). TheVajont Limestone thins basinward and onlaps parts ofthe Trento Platform to the west. Well penetrations in thePo Plain and northern Adriatic Sea (nonproductive) sug-gest that the Vajont sediment is present to the south inthe subsurface (Bosellini et al., 1981; Cati et al., 1987).

Within the Belluno Basin, the Vajont Limestoneoverlies dense, chert-rich micritic limestone and shalebelonging to the Igne Formation (Toarcian–Aalenian;Figures 2, 3, 5, and 6). The Fonzaso Formation(Callovian–Lower Kimmeridgian?) overlies theVajont Limestone and contains cherty, skeletal-richturbidites and debris flows. In the central BellunoBasin, the Fonzaso Formation grades upward into nodu-lar, micritic red limestone belonging to the UpperAmmonitico Rosso (Kimmeridgian–Tithonian), whichin turn grades into thick, hemipelagic white limestone ofthe Biancone Formation (Tithonian–Cretaceous).Toward the east, the Fonzaso Formation andAmmonitico Rosso grade into the Soccher Formation(Lower Kimmeridgian–Cretaceous), which containsresedimented shallow-water carbonate and hemipelagiclimestone. Along the western margin of the Friuli Platform, the Soccher Formation directly overlies theVajont Limestone (e.g., Mt. Sestier section; Zempolich,1995) and passes upward from thin-beddedpeloidal/skeletal grainstone to thick skeletal-rich bedsto massive coral and Ellipsactinia (hydrozoan) reefs andback-reef Nerinacea gastropod grainstone (CellinaLimestone; Upper Oxfordian–Lower Kimmeridgian).The progradation of Upper Jurassic slope and reef

margin sediments over the Vajont area along the easternBelluno Basin indicates that Vajont sediments foundhere and to the west were deposited in periplatform,slope, and basinal settings (Zempolich, 1995). Along thewestern edge of the Belluno Basin, slope and basinalsediments of the Vajont limestone and Fonzaso Forma-tion onlap downfaulted blocks and margins of the eastTrento Platform (Bosellini et al., 1981).

DOLOMITE FIELD RELATIONSHIPS

Regional and Stratigraphic Distribution of Dolomite

Regional field mapping has established that dolomi-tization is mostly confined to slope and basinal facies ofJurassic and Cretaceous sediments in the central andwestern Belluno Basin (Zempolich, 1991a, b; 1995).

Toward the east, the source of the resedimentedooids, only Vajont limestone is found in periplatformareas adjacent to the western margin of the Friuli Plat-form (e.g., Mt. Sestier); dolomite bodies are conspicu-ously absent (Figures 2, 3, 5, and 6). Toward the west,dolomite first occurs in the central Belluno Basin at theVajont Dam and Col Visentin localities (Figures 2, 3, 5,and 6). At the Vajont Dam, dolomitization of slope andbasin facies has resulted in formation of an isolated,rootless dolomite plume ≥300 m high and ~100–200 mwide. At Col Visentin, dolomitization occurs along smallfaults and fractures within a predominantly limestonesection. Other occurrences of Vajont dolomite within the

Figure 1. Early and MiddleJurassic paleogeography ofthe Venetian Alps (modifiedfrom Bosellini et al., 1981;Cati et al., 1987). The studyarea is divided into severalcarbonate platform andbasin domains, includingthe Trento Platform,Belluno Basin, FriuliPlatform, and Tolmin Basin.The Friuli Platform is subdi-vided into several platformsand basins based on theinterpretation of seismicreflection data from theFriuli Plain and Po Basin(Cati et al., 1987).

Page 144: Reservoir Quality Prediction in Sand and Carbonates

central Belluno Basin are found: (1) along the intersec-tion of the Piave Graben and the mouth of the VajontCanyon; (2) in the subsurface of the Piave Graben (i.e.,the Belluno 1 well); and (3) at Villanova. The total extentand geometry of these last three bodies is poorly knowndue to limited exposure and data.

Within the western Belluno Basin, dolomite bodiesspan the entire Jurassic basinal succession and climbupward into the lower Cretaceous section (Figures 2, 3,5, and 6). At Val Zoldo, an isolated, rootless dolomiteplume ≥300 m high and ~100 m wide penetrates theSoverzene, Igne, and Vajont Limestone formations.At the San Boldo, Col dei Moi, and Val Sassuma sec-tions, massive dolomitization (thickness >400 m over20–25 km) has affected the Igne, Vajont Limestone, andFonzaso formations that are now exposed in the crest ofthe Mt. Grappa–Visentin anticline. At Val Sassuma andMt. Tomatico, dolomitization continues higher in thestratigraphic section along vertically oriented paleo-synsedimentary breccia, and locally replaces theFonzaso, Ammonitico Rosso, and Biancone formations.

Massive dolomite is also found at the boundarybetween the western Belluno Basin and the easternedge of the South Trento Platform, where an abrupttransition takes place from limestone platform facies to

a thin belt of dolomitized platform facies to completelydolomitized basinal facies (Figures 2, 3, and 5). At theplatform margin, meter-scale occurrences of replace-ment dolomite are found along fractures and faults thatpenetrate platform strata (e.g., Upper Pliensbachianreef sediments, Mt. Grappa; Zempolich 1993, 1995).Platform strata associated with other large structuralfeatures such as the Seren Graben and other north-south–trending paleolineaments (Figures 2, 3) that com-prise the eastern platform/basin boundary fault arealso massively dolomitized, making stratigraphic corre-lations difficult (e.g., Dolomie Selcifere, Calcari Grigi,Toarcian–Aalenian) (Masetti, 1971; Trevisiani, 1991).

Structural and Crosscutting Relationships of the Dolomite Bodies

Regional and local detailed field mapping ofdolomite bodies within Jurassic basinal sediments of theBelluno Basin indicates that dolomite bodies are linkedto fracture and fault systems associated with Alpinedeformation that was imposed on the Southern Alpsduring the Late Paleogene to Neogene (Figures 3, 7, and8). The dolomite bodies include: (1) an extensive wedge-shaped body (~20–25 km long, 10–15 km wide, and400–500 m thick) that has replaced the Vajont and other

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 131

Figure 2. Stratigraphic agerelationships and distribu-tion of dolomite within theMesozoic area of theVenetian Alps (stratigraphymodified from Bosellini etal., 1981). Massive replace-ment dolomitization is pre-dominantly confined to theMiddle Jurassic VajontLimestone in the central andwestern portions of theBelluno Basin. Dolomitealso occurs as plume-shapedbodies within the underly-ing Lower JurassicSoverzene and Igne forma-tions, and as dolomite “tow-ers” in overlying UpperJurassic–Lower Cretaceousstrata following paleosyn-sedimentary dikes.Dolomitization has alsoaffected the eastern marginof the South TrentoPlatform. These crosscuttingrelationships indicate thatdolomitization occurredduring or following theEarly Cretaceous.

Page 145: Reservoir Quality Prediction in Sand and Carbonates

132 Zempolich and Hardie

basinal sediments present in the core of the Mt.Grappa–Visentin anticline (Figures 9, 10); (2) large,“rootless” dolomite plumes (hundreds of meters thickand high) that have penetrated upward into the VajontLimestone from underlying Lower Jurassic strata (Fig-ures 11–13); (3) large cylindrical and elliptical shapedplumes that penetrate upward from dolomitized VajontLimestone through Upper Jurassic and Lower Creta-ceous strata along Cretaceous-age synsedimentary brec-cias (Figure 14); and (4) smaller meter-scale dolomitebodies found along faults and fractures (Figure 15).

DOLOMITE BODIES AND REACTION FRONTS

Within the Belluno Basin, both large-scale (kilometer-scale) and small-scale (meters to hundreds of meters)dolomite bodies are present. Dolomitization fronts(Wilson et al., 1990) are noted in outcrop by an easilyrecognized and distinct transition from browndolomite to blue limestone. In the southern study area(i.e., the massive dolomite wedge located within thehanging wall of the Mt. Grappa–Visentin anticline),transitions from partially dolomitized to completely

dolomitized limestone occur over distances of severaltens of centimeters to hundreds of meters. In thenorthern study area, where small-scale dolomiteplumes and fault-related dolomite bodies penetrateupward through the stratigraphic section, the dolomi-tization fronts are relatively sharp and occur over dis-tances of centimeters to tens of centimeters.

Large-Scale Dolomite Bodies

Areal Distribution

In the southern study area, a massive areally exten-sive wedge of replacement dolomite is found within theVajont Limestone and other basinal sediments on thehanging wall of the Bassano Line (Figures 3, 5, 7–10).The hanging wall is the southward-dipping limb of theM. Grappa–Visentin anticline, which trends N60–80°E.This dolomite body is >400 m thick at Passo di San Boldoand Col dei Moi and extends laterally 20–25 km, paral-leling the overthrust from northeast to southwest. To thenortheast, where the anticline wraps around toward thenorth-northeast, the dolomite exposures are reduced toseveral small occurrences (meter-scale) at Col Visentin.To the southwest, the Vajont limestone is completely

Figure 3. Tectonic map of the southern Alps (modified from Doglioni, 1990). Mesozoic platform and basin stratawere thrusted toward the south during the Alpine orogeny (Late Oligocene to Recent) (Massari et al., 1986;Doglioni, 1990). Study localities are marked by circled letters. Southern study localities include: B = Passo SanBoldo; C = Col Visentin; D = Col dei Moi; E = Grigno; F = Fontana Secca; G = Mt. Grappa; L = Valpore di Cima;M = Mezzamonte; P = Ponte Serra; T = Mt. Tomatico; U = Val Sassuma. Northern study localities include: A =Villanova; S = Mt. Sestier; V = Vajont Dam/Canyon; W = Belluno 1 exploration well; Z = Val Zoldo/Igne.Geographic distribution of dolomite bodies is noted by dolomite shade pattern. Most dolomite bodies occurwithin Jurassic and Cretaceous basinal sediments of the Belluno Basin. A major dolomite body, ~25 km ×15 kmin area and ≥400 m in thickness, is present in the southern study area and is located within the crest of the Mt.Grappa–Visentin anticline. Isolated plume-shaped dolomite bodies, which are ~200–300 m in width and ≥300 min height, occur in the northern study area and are hosted within footwall synclines of the Belluno thrust sheet.

Page 146: Reservoir Quality Prediction in Sand and Carbonates

dolomitized as far as the border fault associated with theeastern margin of the South Trento Platform (Figure 3).Mapping to the north of this front (Quero-Vas locality)suggests that the dolomite front wedges out over10–15 km to the north-northwest, with the wedge thin-ning downward through the stratigraphic section.

In the central part of the Mt. Grappa–Visentin anti-cline, the dolomite front penetrates the crest of theanticline and passes upward through the Vajont Lime-stone and into the overlying Fonzaso Formation atPasso di San Boldo (Figures 9, 10). At Passo di SanBoldo, a flower structure is recognized in the crest of theanticline by the presence of large-scale vertical to sub-vertical faults that were formed during regional trans-pression (Doglioni, 1990). These faults and relatedfractures are reflected in the present-day topography assmall canyons oriented parallel to the crest of the anti-cline (Figures 9, 10). It is in these fault-controlledcanyons that the dolomite fronts can be observed pene-trating upward into the Fonzaso Formation. Elsewhere,the upper dolomitization front is typically found nearthe top of the Vajont Limestone or higher, where thestratigraphy is transected by vertically oriented faults orbreccia (e.g., paleosynsedimentary breccia in UpperJurassic and Lower Cretaceous strata; Val Sassuma andMt. Tomatico). Offset of the stratigraphic section alongthese faults is minor, and dolomitization fronts can beobserved at the top of the section where remnants ofunaltered Vajont limestone are preserved in the eastwall of the San Boldo Canyon. The Vajont Limestone isentirely dolomitized from Passo di San Boldo to Col dei

Moi and out into the leading edge of the anticline, whereit disappears into the subsurface. Faulting apparentlyhas controlled dolomitization, because lateral contactsbetween dolomite and limestone are commonly abrupt.Passing northward away from the crest of the anticlineand into the Belluno thrust sheet, the limestone–dolomite contact stratigraphically drops within theupper 100 m of the Vajont limestone. This upper contactcan be viewed in both the east and west hills on eitherside of the pass at Col dei Moi. At the base of Col deiMoi, limited exposure of the Vajont–Igne contact indi-cates that dolomitization has also affected the upper-most portion of the underlying Igne Formation. Fromhere, the dolomite body disappears downward into thesubsurface. Total thickness of the dolomite body alongthe Mt. Grappa–Visentin anticline may exceed400–500 m where dolomitization of the underlying Igneand Soverzene formations has occurred.

Continuing toward the southwest along the Mt.Grappa–Visentin anticline, the Vajont Limestone is mas-sively dolomitized in the vicinity of the eastern marginof the South Trento Platform (Figures 2, 3, and 5). In thisregion, both platform and basinal strata are complexlyfaulted due to Tertiary uplift. Also found in associationwith massive Vajont dolomite are isolated dolomite andlimestone breccia bodies that penetrate upwardthrough the Fonzaso, Ammonitico Rosso, and Bianconeformations at the Val Sassuma and Mt. Tomaticolocalities (Figures 2, 5, and 14). These vertically orienteddolomite breccias are ellipsoidal in shape, penetrate100–200 m into the Upper Jurassic–Lower Cretaceous

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 133

Figure 4. Distribution of theVajont Limestone based onboth outcrop and subsurfacedata (modified fromBosellini et al., 1981; Cati etal., 1987). The Vajont lime-stone thickens toward theFriuli Platform and attains amaximum thickness of≤600 m near the platformmargin. Toward the west, theVajont limestone thins andonlaps portions of the TrentoPlatform.

Page 147: Reservoir Quality Prediction in Sand and Carbonates

134 Zempolich and Hardie

Figu

re 5

. Pal

eoge

ogra

ph

ic d

epos

itio

nal

pro

file

an

d s

trat

igra

ph

ic c

ross

sec

tion

sh

owin

g d

istr

ibu

tion

of

dol

omit

e w

ith

in th

e d

omi-

nan

tly

lim

esto

ne

Ven

etia

n A

lps

(sou

ther

n s

tud

y ar

ea).

Ou

tcro

p lo

cali

ties

, in

clu

din

g sa

mp

lin

g tr

aver

ses

and

mea

sure

d s

ecti

ons,

are

not

ed b

y b

lack

bar

s (F

igu

re 3

). M

assi

ve r

epla

cem

ent d

olom

itiz

atio

n is

pre

dom

inan

tly

con

fin

ed to

slo

pe

and

bas

in f

acie

s of

the

Vaj

ont L

imes

ton

e, b

eau

tifu

lly

exp

osed

in th

e co

re o

f th

e M

t. G

rap

pa–

Vis

enti

n a

nti

clin

e.

Page 148: Reservoir Quality Prediction in Sand and Carbonates

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 135

Figu

re 6

. Pal

eoge

ogra

ph

ic d

epos

itio

nal

pro

file

an

d s

trat

igra

ph

ic c

ross

sec

tion

sh

owin

g d

istr

ibu

tion

of

dol

omit

e w

ith

in th

e d

omi-

nan

tly

lim

esto

ne

Ven

etia

n A

lps

(nor

ther

n s

tud

y ar

ea).

Ou

tcro

p lo

cali

ties

, in

clu

din

g sa

mp

lin

g tr

aver

ses

and

mea

sure

d s

ecti

ons,

are

not

ed b

y b

lack

bar

s (F

igu

re 3

). R

epla

cem

ent d

olom

itiz

atio

n is

con

fin

ed to

dol

omit

e p

lum

es th

at a

re h

oste

d w

ith

in s

lop

e an

d b

asin

faci

es o

f th

e S

over

zen

e, I

gne,

an

d V

ajon

t L

imes

ton

e fo

rmat

ion

s.

Page 149: Reservoir Quality Prediction in Sand and Carbonates

136 Zempolich and Hardie

section, and pass upward into undolomitized limestonebreccia. The original brecciation of Jurassic–LowerCretaceous limestone is interpreted by Doglioni (1990)and Masetti (1990, personal communication) to haveoccurred during formation of Cretaceous-age synsedi-mentary dikes. The presence of relic limestone brecciaabove dolomitized breccia indicates that the dolomitiz-ing fluids originated from below the Upper Jurassic toCretaceous section and ascended along the breccia,which was more permeable than surrounding bedded,chert-rich micritic limestone.

To the northeast, the massive dolomite body foundin the core of the Mt. Grappa–Visentin anticline rapidlythins to several-meter-thick occurrences of dolomite atCol Visentin. Here, small-scale dolomite reaction frontsare found in association with minor faults (Figures 3,5). Calculated volumes of this massive dolomite wedgerange from 50 to 94 km3.

Isolated Dolomite Plumes

Areal DistributionIn the northern study area, several isolated dolomite

bodies hundreds of meters in height and width arefound at Vajont Canyon and Val Zoldo. These dolomite

bodies are located on the trailing edge of the Bellunothrust sheet in or near the axis of an east-west–orientedfootwall syncline (Figures 3, 7, and 8). These isolatedplume-shaped bodies are oriented upward through thestratigraphic section and are cored by hydrothermalbreccia. The occurrence of a succession of such isolatedplume-shaped bodies along the same structural trendsuggests that the dolomite bodies were formed by flowof Mg-bearing fluid along the axis of the east-west foot-wall syncline.

Vajont Canyon

Remnants of a massive plume-shaped dolomitebody, >300 m high and several hundred meters wide,are found on both the north and south walls of theVajont Canyon just to the west of the Vajont Dam (Fig-ures 11, 12). The dolomite body is distinguished by adistinct color change from brown (dolomite) to blue(limestone) in both the north and south canyon walls.Along the north canyon wall, the vertically orienteddolomite body is discordant, with bedded limestonelying at low-angle dip, and has the shape of a simpleupward-oriented plume (Figure 11). Discordance is

Figure 7. Distribution of dolomite within the present-day thrust and fold structural configuration of theVenetian Alps (structural interpretation modified from Doglioni, 1990). Dolomite bodies include a majorwedge of dolomite located within the crest and hanging wall of the Mt. Grappa–Visentin anticline. Isolateddolomite plumes are found in footwall synclines located along the trailing edge of the Belluno thrust sheet.Movement of the thrusts can be accurately dated by analysis of sedimentation events and patterns in theVenetian foredeep (Massari et al., 1986; Doglioni, 1990). These data indicate that thrusting was initiated by thelate Oligocene. Doglioni (1990) suggests that some thrusting may have begun even earlier, as indicated byonlap relationships of early Eocene sediment. Extensive uplift and subaerial exposure of the growing fold beltbegan during the early middle Miocene and continued into the Pliocene. T = Late Permian–Middle Triassic, B =Early Cretaceous (Biancone Formation), C = crystalline basement, E = Paleogene, J = Jurassic, N = Neogene, P =Late Triassic (Dolomia Principale), Q = Quaternary, S = Late Cretaceous (Scaglia Rosso Fm.).

Page 150: Reservoir Quality Prediction in Sand and Carbonates

characterized by the abrupt disappearance of limestonebedding planes at the contacts between limestone anddolomite on each side of the plume. The presence ofbedded limestone over the top of the plume marks theupper limits of the dolomite–limestone contact alongthe north canyon wall. At the base of the south canyonwall, the vertical margin of the dolomite plume issharply discordant, with bedded limestone now lyingat low-angle dip (Figure 12). Moving upward from thebase, the dolomite body turns toward the west, becom-ing concordant with bedded limestone and eventuallypinching out. Thus, the top of the dolomite body alongboth the north and south walls of the Vajont Canyon

appears to be confined to the upper Vajont limestone.Along the Vajont River, at the base of both north andsouth canyon walls, the dolomite plume disappearsinto the subsurface.

At the edges of the main dolomite plume, wedge-shaped apophyses of dolomite and associated limestone–dolomite transitions emanate from the main dolomitebody and follow bedding planes and fractures into sur-rounding unaltered limestone (Zempolich, 1995).Dolomite–limestone transitions are narrow bands thatrange in thickness from several centimeters to severalmeters. At the center of the dolomite plume, the replace-ment dolomite is extensively brecciated and cemented

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 137

Figure 8. Geographic distribution of dolomite within anticlines and synclines of the Venetian Alps (structur-al interpretation modified from Massari et al., 1986; Doglioni, 1990). Major dolomitization is related to faultand fracture systems that are associated with anticlines and synclines formed during Tertiary compression(figure numbers refer to exemplary photographic plates or sketches of dolomite bodies). The major dolomitefront located within the crest and hanging wall of the Mt. Grappa–Visentin anticline is ~25 km long, 10–15 kmwide, and ≥400 m thick, and represents 50–94 km3 of dolomite. Toward the northeast, the dolomite front thinsand occurs within several meter-scale beds of Vajont limestone. In the central portions of the Mt. Grappa–Visentin anticline, the dolomite front climbs upward through the stratigraphic section. Toward the south-west, the dolomite front broadens and affects the eastern margin of the South Trento Platform. To the north,isolated dolomite plumes are located within footwall synclines of the Belluno thrust sheet. Each of thesebodies represents ~2.4 ×10–2 km3 of dolomite.

Page 151: Reservoir Quality Prediction in Sand and Carbonates

138 Zempolich and Hardie

by thick linings of baroque dolospar cement, whichcompletely fills interclast pore space. The angular natureof the dolomite clasts indicates that replacement dolomi-tization preceded brecciation and baroque dolo-mite cementation. Moreover, breccia and baroquedolomite cement are only found within the interior ofthe dolomite plume. These petrographic relationshipsdemonstrate that brecciation and baroque dolomitecementation were the final diagenetic events associatedwith replacement dolomitization (Zempolich, 1995).

Dolomite is also found at the mouth of the VajontCanyon (east wall of the Piave Graben), located ~0.5 kmto the west of the main dolomite body just described.While similar replacement dolomite and breccia arepresent at this locality, the full extent of this dolomitebody is unknown due to normal faulting during the lateTertiary, and downdropping of the stratigraphic sec-tion into the Piave Graben. However, this dolomite,together with small occurrences of replacementdolomite outcropping along the west wall of the Piave

Figure 9. Dolomite in out-crop at Passo di San Boldo,Mt. Grappa–Visentin anti-cline. Top photograph:View looking northwardinto the crest of the Mt.Grappa–Visentin anticline.Tunnels and road climbthrough cliffs (≥400 m)composed of Vajontdolomite. This outcrop ispart of an extensive wedgeof dolomite that is hostedwithin the crest and hang-ing wall of the anticline(see Figures 3 and 7). Lowerinset: Topographic map ofthe Passo di San Boldo area(5-m contour interval).Large-scale faults and frac-tures (dashed lines; Figure10) are oriented parallel tothe axis of the anticline,which trends N60°E. Thedolomite front climbsupward along faults andfractures through the strati-graphic section at this local-ity; it has also affected theUpper Jurassic FonzasoFormation. Present-daydrainages accentuate thefaults and fractures, whichminimally offset the strati-graphic section ~10 m).

Page 152: Reservoir Quality Prediction in Sand and Carbonates

Graben (e.g., at Villanova and Longarone), suggests thepresence of a second dolomite body (plume?), much ofwhich might be buried beneath the Piave Graben. Theoccurrence of dolomitized Vajont limestone in the PiaveGraben is confirmed by an exploration well (AGIP Bel-luno 1 well), located just to the south of these outcrops,that penetrated and cored the Vajont Limestone directlybeneath the Dolomia Principale Formation, which formsthe hanging wall of a buried overthrust (AGIP, personalcommunication, unpublished well results). Exposure ofVajont dolomite along graben walls and within buriedthrusts of the Piave Graben indicates that dolomitizationof the Vajont occurred before down-faulting of segmentsof the Belluno and Moline thrust sheets (Figures 3, 7)into the Piave Graben during the Late Tertiary.Val del Zoldo/Igne

The upper section of a large dolomite plume is foundwithin the chert-rich micritic Soverzene and Igne forma-tions near the town of Soffranco (Figure 13). This plumeis located within the same footwall synclinorium as theplume described at Vajont Canyon, but is on the oppo-site side of the Piave Graben. Large clasts of dolomitizedSoverzene and Igne carbonate, fractured chert clasts,dolomitized geopetal silt, and baroque dolospar charac-terize the hydrothermal breccia found in the center ofthis body (Zempolich, 1995). The brecciated core of thisbody is ~100 m wide and ≥200 m high. Along the edgesof the body, thin “fingers” of dolomite breccia (tens ofcentimeters to several meters thick) follow bedding

planes for ≤10 m before grading into cherty argillaceousmicrite. Angular dolomite clasts and the presence ofdolomitized geopetal silt indicate that replacementdolomitization preceded and overlapped brecciation,and preceded the precipitation of baroque dolomitecement. These petrographic relationships are similar tothose found in the breccia at the Vajont Dam locality andcarry the same implication; that is, the brecciation andbaroque dolomite cementation were the final diageneticevents associated with replacement dolomitization.

Directly above the main body of the dolomite plume,replacement dolomitization can be followed upwardfrom the brecciated dolomite core along small fracturesand faults (Figure 13). Replacement dolomitization con-tinues along these pathways stratigraphically upwardthrough the Soverzene and Igne formations and into theoverlying Vajont Limestone. A replacement dolomitehalo is present within nonbrecciated Vajont lithologiesexposed in the cliff above the plume and in nearby out-crops lacking breccia located behind the cliff along trailsleading west from the town of Igne. Examination of lim-ited exposures of the Fonzaso and Ammonitico Rossoformations suggests that they were not affected bydolomitization at this locality. Near the town of Igne,undolomitized Igne and Vajont bedded limestone isexposed. These field relationships indicate that replace-ment dolomitization was restricted to the plume-shaped dolomite breccia body found within theSoverzene, Igne, and Vajont strata at Soffranco, and

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 139

Figure 10. Vertical to subvertical faults within the crest of the Mt. Grappa–Visentin anticline at Passo di SanBoldo (view is toward the west wall of the pass; Figure 9). Arrows point to large faults and fractures that dissectcliffs composed of Vajont dolomite (≥400 m). Faults and fractures are oriented parallel to the anticline, whichtrends N60°E. The dolomitization front continues along the south limb of the anticline, where it disappears intothe subsurface and forms the hanging wall of the Bassano thrust (left part of photograph; Figures 3 and 7).

Page 153: Reservoir Quality Prediction in Sand and Carbonates

140 Zempolich and Hardie

(A)

(B)

Page 154: Reservoir Quality Prediction in Sand and Carbonates

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 141

suggest that dolomitizing fluids ascended from depth.Reaction fronts and textural transitions between initialreplacement dolomite at Val del Zoldo/Igne are similarto those observed at the Vajont Dam.

Small-Scale Dolomite Occurrences

Meter- to decameter-scale dolomite bodies occur asisolated reaction fronts, or occur in association withlarge-scale dolomite bodies. These small-scaledolomite bodies provide important petrographic andgeochemical evidence of fluid movement, the mecha-nism by which precursor limestone was replaced, andthe formation of pore space through dolomitization(Zempolich, 1995). Small-scale dolomite bodies occuras: (1) small wedges (10–30 m) found parallel and sub-parallel to fault, fracture, and bedding planes; and (2)strata-bound beds. Small-scale reaction frontsbetween dolomite and limestone are seen in outcropas sharp fronts, on the scale of centimeters or less, andtransitional fronts over a distance of several meters,grading from zones of completely dolomitized rock topartially dolomitized limestone to unaltered lime-stone. Detailed analysis of closely spaced samplesacross these fronts indicates that Vajont dolomite tex-tures and compositions show progressive texturaland compositional maturity with increasing proxim-ity to fluid conduits (Kupecz and Land, 1994; Zem-polich, 1995). A particularly well defined example of atransitional front is found at Villanova. Here, a smalldolomite body ~20–30 m wide is exposed along a roadcut that dissects a small rollover anticline in theVajont limestone. Dolomite-to-limestone transitionsat this locality occur along a small fault that cutsobliquely across the bedding and within bedded lime-stone. The dolomite front on the south side of thebody is wedge-shaped, narrowing upward along thefault plane (Figure 15), pinching out obliquelybeneath undolomitized bedded limestone where thefault soles out into a bedding plane. The front on thenorth side of the body is a simple gradation from par-tially dolomitized limestone to unaltered limestonewithin an individual bed, and has been described indetail by Zempolich (1995).

STRATIGRAPHIC CONSTRAINTS ONTHE TIMING OF DOLOMITIZATION

The occurrence of crosscutting relationships betweendolomite bodies and the host Jurassic and Cretaceousstrata places constraint on the timing of dolomitization.These data indicate that dolomitization occurred duringor following the Early Cretaceous and was focusedalong structural features related to paleolineaments andTertiary-age deformation.

Lower Jurassic Dolomite Bodies

Dolomite bodies hosted in the Lower JurassicSoverzene and Igne formations occur as “rootless”plumes, and pass into dolomite found in the overlyingVajont Limestone. At the Val del Zoldo locality (Fig-ures 2, 6, and 13), the large, partly brecciated dolomitebody penetrates upward through cherty, dark micriticlimestones of the Soverzene and Igne formations andinto the overlying Vajont Limestone. Some brecciationof Soverzene limestone is attributed to the occurrenceof synsedimentary (Lower Jurassic) growth faults andslumps, examples of which are widespread in thewestern Belluno Basin and Alpi Feltrine (Masetti andBianchin, 1987). Dolomitization of the Igne Formationis also observed in the hanging wall of the Mt.Grappa–Visentin anticline and underlies a thick occur-rence of dolomitized Vajont limestone (Figures 2, 3, 5,and 7). At these localities, crosscutting relationships ofthe Lower and Middle Jurassic dolomite bodies indi-cate that dolomitization must have occurred during orfollowing the Middle Jurassic.

Upper Jurassic Dolomite Bodies

Bodies of dolomitized Vajont limestone located infault zones cut stratigraphically upward into the overly-ing Upper Jurassic section [e.g., the Mt. Grappa–Visentin anticline (Figures 2, 3, 5, and 7)]. Such dolomitecrosscutting relationships indicate that dolomitization ofthe Vajont Limestone must have occurred during or following the Late Jurassic. Furthermore, the largedolomite wedge of dolomitized Vajont limestone associ-ated with crestal faults within the Mt. Grappa–Visentinanticline suggests that dolomitization may be related toTertiary deformation and the formation of the VenetianAlps thrust belt during the late Oligocene–Recent.

Lower Cretaceous Synsedimentary Breccia

At Val Sassuma and Mt. Tomatico, massive Vajontdolomites can be traced upward into the Fonzaso,Upper Ammonitico Rosso, and Biancone formations(Figures 2, 6, and 14). These vertically oriented brecciasare roughly columnar in shape, penetrate 100–200 minto the Upper Jurassic–Lower Cretaceous section, andpass upward into limestone breccia. The original brec-ciation of Jurassic–Lower Cretaceous limestone isinterpreted by Doglioni (1990) to have occurred dur-ing formation of Cretaceous-age synsedimentary dikesin association with extensional tectonics. At these

Figure 11. (A) Large dolomite plume exposed alongthe north wall of the Vajont Canyon (tunnel locatedon the right side of dolomite body is ~8 m high;downward and upward dimensions of this photomontage are distorted by the camera angle). Dolomite(DOL) is dark brown in outcrop and crosscuts beddedlimestone (arrows), which is light blue (LS). Theplume is 200–300 m wide and >300 m high. It isbounded above by bedded limestone and disappearsbelow into the subsurface. (B) Schematic of (A)depicting the large dolomite plume exposed alongthe north wall of the Vajont Canyon. Dolomite instipple pattern. Tunnel located on the right side ofdolomite body is ~8 m high.

Page 155: Reservoir Quality Prediction in Sand and Carbonates

142 Zempolich and Hardie

localities, dolomite crosscutting stratigraphic relation-ships indicate that dolomitization of these synsedi-mentary breccias must have occurred during orfollowing the Early Cretaceous.

Collectively, crosscutting stratigraphic relationshipsof dolomite bodies observed throughout the study area

indicate that massive dolomitization of Vajont and otherJurassic and Lower Cretaceous basinal strata occurredduring or following the Early Cretaceous. To provide afurther constraint on the timing of dolomitization, therelationship between Vajont dolomite bodies and Ter-tiary structural elements is examined below.

(A)

(B)

Figure 12. (A) Large dolomiteplume exposed along thesouth wall of the VajontCanyon (the road that is visi-ble crossing the dolomitebody is ~10 m wide; down-ward and upward dimen-sions of this photo montageare distorted by the cameraangle). Dolomite (DOL) isdark brown in outcrop andcrosscuts bedded limestone(dashed lines), which is lightblue (LS). The dolomiteplume is cored by hydrother-mal breccia (Br), which iscomposed of replacementdolomite clasts and baroquedolomite cement. The plumeis 200–300 m wide and >300 mhigh. Toward the top, thedolomite body becomes con-cordant with bedded lime-stone and eventually pinchesout. At the base, the bodydisappears below into thesubsurface. Numbers refer todetailed sampling that wasconducted along the damaccess road, which crosses theplume (Zempolich, 1995).Arrows point in the directionof probable fluid flow duringdolomitization. (B) Schematicof the large dolomite plumeexposed along the south wallof the Vajont Canyon.Dolomite in stipple pattern;breccia core noted by clastpattern. The road that is visi-ble crossing the dolomitebody is ~10 m wide.

Page 156: Reservoir Quality Prediction in Sand and Carbonates

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 143

RELATIONSHIP BETWEEN DOLOMITEBODIES AND ALPINE DEFORMATION

Dolomite bodies distributed within Mesozoic basi-nal sediments are exposed within hanging-wall anti-clines and footwall synclines that were formed duringlate Oligocene–Recent thrusting (Figures 3, 7, and 8).The specific distribution of dolomite along these struc-tural features suggests that extensional and transpres-sive faulting in the axes of these anticlines andsynclines controlled the circulation of dolomitizingfluid. If this is true, dolomitization of Mesozoic basinalsediments must have occurred sometime between theinitiation of thrusting and regional epierogenic uplift(i.e., subaerial exposure) of the sequence.

The timing of initial thrusting and the uplift history inthe southern Alps has been accurately dated by studiesof pre- and synorogenic sediments that were depositedand shed into the surrounding foredeeps. Deformationof the southern Alps began during the Late Cretaceous,and progressed from to east to west due to the obliqueconvergence of Europe and Adria during the Tertiary(Massari et al., 1986). To the west of the study area, initialdeformation is recorded by the deposition of Late Creta-ceous and Early Tertiary flysch in both the Lombard andwestern Venetian basins (Gaetani and Jadoul, 1979; Mas-sari et al., 1986). To the east, thrusting related to lateEocene compression (Dinaric fold belt) deposited flyschin the eastern Venetian basin (Doglioni, 1990). Majordeformation of the study area (i.e., the central VenetianAlps) occurred during the later phases of regional trans-pressive deformation (Late Oligocene–Recent) and ischaracterized by a series of south-vergent thrusts (trend-ing N60–80°E) that involve crystalline basement to thenorth (Doglioni, 1990). Major thrusts, from south tonorth, include the Bassano-Maniago, Tezze, Belluno,and Valsugana (Figure 3). However, onlap relationshipsand angular unconformities between early Eocene flysch and late Oligocene molasse in the Belluno syn-cline (i.e., the trailing edge of the Mt. Grappa–Visentinanticline) suggest that initial detachment and thrustingmay have begun earlier (Doglioni, 1990).

Prior to major thrusting and development of theVenetian foredeep, marine siliciclastic and carbonateshelf sediments of late Oligocene to early middleMiocene age were deposited regionally across the studyarea under the influence of the Dinaric fold belt to theeast (Chattian to Langhian cycle; Massari et al., 1986).Major thrust movement and loading in the study areabegan at least by early middle Miocene time (Ser-ravalian) and led to foreland subsidence and thick accu-mulation of hemipelagic marls and mudstones, anintermediate basin-fill sequence, and fan-delta and allu-vial deposits (Serravalian to Recent cycle; Massari et al.,1986). This synorogenic sedimentary sequence marks achange in the polarity of sedimentation of the VenetianBasin and records the first major movement of thrustsheets toward the south. Uplift and denudation of thesethrust sheets in the Late Tertiary can be accuratelydated by the inclination and alteration of molasse sedi-mentation patterns, which were shed off of the growinganticlines. Progressive inclination of sediment packages

and formation of angular unconformities occur alongboth dip and strike sections (Massari et al., 1986;Doglioni, 1990). These data indicate that major thrust-ing and uplift took place rapidly from the early middleMiocene to the late Pliocene (~10 Ma).

The spatial association of Vajont dolomite with theseTertiary thrust features (Figures 3, 7, and 8) suggeststhat dolomitization of Mesozoic basinal sediments mayhave occurred within these thrust sheets sometimebetween the early Eocene/late Oligocene and the lateOligocene/early middle Miocene (in agreement withstratigraphic crosscutting relationships exhibited bydolomite bodies, which indicates that dolomitizationpostdated the Lower Cretaceous). During this timerange, the study area, and specifically the Mesozoicbasinal succession, was still buried beneath several kilo-meters of section and was located beneath coastal andmarine environments. The entire region experiencedmajor uplift during the early middle Miocene to latePliocene. If uplift and surficial expression of the Mt.Grappa–Visentin anticline and other compressionalstructures occurred by the middle Miocene (Massari etal., 1986), structural constraints place the timing ofdolomitization somewhere between the initiation ofcompressional tectonics during the early Eocene, andthe uplift and exposure of the advancing thrust sheetsduring the early middle Miocene. Dolomitization musthave been completed prior to extensive uplift because(1) significant topographic expression of the anticlinewould have initiated meteoric recharge, and theoreti-cally would have shut down the subsurface circulationof Mg-bearing fluid; and (2) metastable dolomitereplacement textures and fronts are beautifully pre-served in these outcrops, which suggests that dolomiti-zation preceded uplift.

In summary, field and stratigraphic relationshipsindicate that dolomitization of Mesozoic-age sedi-ments in the Venetian Alps is mostly confined to slopeand basin facies contained in the Belluno Basin. Withinthese basinal strata, the majority of dolomite occurs asmassive and isolated bodies within the Vajont Lime-stone and as isolated bodies beneath and above Vajontdolomite. Structural and stratigraphic crosscuttingrelationships collectively suggest that late-stagedolomite bodies within the Vajont and other basinalsequences were formed following the Lower Creta-ceous. The spatial distribution of these dolomite bod-ies within otherwise tight basinal strata and theirrelationship to Tertiary-aged compressional struc-tures, other paleolineaments, and (paleo)synsedimen-tary faults suggest that dolomitizing fluids werefocused along zones of structurally enhanced porosityand permeability during Tertiary deformation. Thetiming of these Alpine structural events indicates thatdolomitization occurred sometime between the earlyEocene and the early middle Miocene during initialcompression and prior to rapid uplift of the regionduring the early middle Miocene to the Pliocene.

These conclusions about the dolomitization of theVajont and other basinal strata of the Belluno Basin arequite different from those reached by Cervato (1990) forthe dolomite bodies in the nearby Lessini Mountains

Page 157: Reservoir Quality Prediction in Sand and Carbonates

144 Zempolich and Hardie

(A) (B)

Page 158: Reservoir Quality Prediction in Sand and Carbonates

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 145

Figure 14. Dolomitized Vajont limestone and Upper Jurassic–Lower Cretaceous dolomite breccia, Val Sassuma,Mt. Grappa–Visentin anticline. (A) Panorama of the south wall of Val Sassuma depicting the Mesozoic basinalstratigraphic succession composed of the Vajont Limestone (Vj), the Fonzaso Formation (Fz), and the BianconeFormation (B). The Ammonitico Rosso (2–4 m thick) is found just above the Fonzaso Formation. The VajontLimestone, which is 400–500 m thick at this location, is massively dolomitized. The dolomite front is located nearthe contact between the Vajont Limestone and the Fonzaso Formation, except where the front climbs upwardthrough the stratigraphic section (B–D) (Figures 2, 6). In these instances, partial dolomitization of paleo-synsedi-mentary breccia has formed erosion-resistant “towers” (arrows in C, closeup in B and D) [Doglioni (1990) creditsthese synsedimentary breccias as having originally formed during the Cretaceous]. Higher in the stratigraphicsection, these paleosynsedimentary breccias are composed of limestone.

Figure 13. (A) Large dolomite plume exposed along Val Zoldo, across from the village of Soffranco (truck isshown for scale; upward dimensions of this photo montage are distorted by the camera angle). Dolomiteplume is cored by hydrothermal breccia that is composed of clasts of dolomitized carbonate, chert, andbaroque dolomite cement. Dolomite “fingers” protrude from the main dolomite body and penetrate into sur-rounding limestone along select beds and bedding planes. The breccia core is ~100 m wide and ≥200 m high.The plume penetrates upward through the Soverzene (Sz) and Igne (Ig) formations; a replacement dolomitehalo is present within the Vajont Limestone (Vj) in the cliff above. (B) Schematic of (A) depicting the largedolomite plume exposed at Val Zoldo, across from the village of Soffranco. Dolomite in stipple pattern; brec-cia core noted by clast pattern. Dolomite fingers protrude from the main dolomite body and penetrate intosurrounding limestone along select beds and bedding planes. Truck is shown for scale.

Page 159: Reservoir Quality Prediction in Sand and Carbonates

146 Zempolich and Hardie

(southern Trento Platform). In the latter area, wheredolomite is found predominantly in platform strata,Cervato attributes the dolomitization to the hydrother-mal circulation of seawater related to the emplacementof magmatics within the southern Trento Platform dur-ing the Tertiary. The absence of magmatics in the pres-ent study area and the presence of a thick sequence ofplatform limestone (central Trento Platform) betweenTertiary volcanics to the southwest and the BellunoBasin suggest that dolomitization of the Vajont andother basinal limestone was unrelated to the hydrother-mal dolomitization of platform sequences further to thesouthwest (Cervato, 1990).

PETROGRAPHY AND GEOCHEMISTRY

Limestone Components

The undolomitized Vajont ooids typically displaycortices that are composed of radially oriented, small

subequant to bladed calcite crystals (Zempolich, 1995).Radial calcite fabrics, nonluminescence, enriched 13Cand 18O isotopic compositions (average δ13C = +2.13‰and δ18O = –3.12‰), low covariant Sr-Mg contents, lowMn-Fe contents, and an absence of neomorphic texturecollectively suggest that Vajont ooids were originallycomposed of radial low-Mg calcite and underwent lit-tle diagenesis prior to dolomitization (Zempolich,1995). These data suggest that Vajont ooids were rede-posited in the Belluno Basin as relatively pristine, min-eralogically stable low-Mg calcite.

Aragonitic and high-Mg calcite skeletal grains inthe Vajont limestone that were deposited along withradial ooids in gravity flows are now replaced by low-Mg calcite (Zempolich, 1995). These grains exhibit aspectrum of fabric-retentive and fabric-destructiveneomorphic fabrics, and possess enriched 13C and 18Oisotopic compositions similar to radial ooids. This sug-gests that original metastable components werealtered to low-Mg calcite early in the diagenetic his-tory of the limestone.

Intergranular pores in resedimented ooid grainstonewere first cemented by pore-lining, nonluminescent,equant low-Mg calcite cement. Nonluminescent equantcalcite cement occurs as thin isopachous linings inresedimented grainstone, as intraskeletal pore fill, andas isopachous linings in skeletal molds formed throughthe dissolution of original aragonite (Zempolich, 1995).Isotopic compositions of nonluminescent equant calciteare enriched with respect to other calcite cements andfall within the field defined by radial calcitic ooids.Analogous, equant low-Mg cement has been describedin modern slope and basin settings by Schlager andJames (1978). Based on these data, isopachous nonlumi-nescent equant cement is interpreted as an early marineprecipitate in slope settings. Furthermore, its occur-rence in the ooid grainstones as a pore-lining phase inprimary intergranular voids, intraskeletal pore space,and within skeletal molds indicates that precipitationbegan soon after deposition of carbonate in slope set-tings and continued during shallow burial diagenesis.

Late diagenetic calcite fabrics include banded lumi-nescent equant calcite that overlies nonluminescentequant calcite and fills remaining intergranular porespace, coarse luminescent calcite that fills molds ofskeletal grains, and fracture-filling luminescent calcitethat crosscuts all previously described fabrics. Progres-sive depletion in oxygen values from banded lumines-cent calcite to mold-filling luminescent calcite tofracture-filling luminescent calcite suggests progres-sive cementation in a burial environment (Zempolich,1995). Importantly, late calcite cement occluded themajority of intergranular, intragranular, and moldicporosity that remained after early cementation and dis-solution in slope and shallow burial environments.

In summary, early and late diagenesis of Vajontsediments in basinal settings resulted in the formationof a relatively impermeable and mineralogically stable(low-Mg calcite) volume of rock (Zempolich, 1995).

Limestone–Dolomite Reaction Fronts

Dolomite bodies within the Vajont Limestone in boththe southern and northern study areas exhibit a

Figure 15. Small-scale dolomite–limestone reactionfront, Villanova locality. This mineralogic transitionforms the left side of the dolomite wedge observed.Dolomite fronts (brown; DOL) emanate from a fault(F) and propagate (open arrows) toward the left intounaltered limestone (light blue; LS) and toward theright into the core of the dolomite wedge. Arrowsalong the fault point toward the probable directionof fluid flow.

Page 160: Reservoir Quality Prediction in Sand and Carbonates

spectrum of replacement and recrystallized dolomitefabrics and variable Ca-Mg compositions that illustratethe initial nonmimetic replacement of limestone andprogressive stabilization of intermediate dolomitephases (e.g., Kupecz et al., 1993; Kupecz and Land,1994). These textures and compositions are distributedover centimeter- to meter-scale transitions from partiallydolomitized limestone to completely dolomitized lime-stone and exhibit a concomitant increase in the degree ofneomorphism and recrystallization with increasingproximity to fluid conduits (i.e., fractures, faults, andbedding planes). The inherent metastability and evolu-tion of these initial and intermediate dolomite fabrics, asdefined by petrographic and geochemical study, hasbeen explored by Zempolich (1995).

Transitions from limestone to dolomite occur overseveral tens of centimeters to tens and hundreds ofmeters in relationship to faults and fractures (Figure15). Macroscale replacement fabrics include the grossretainment of lithoclastic grains and sedimentary struc-tures through variations in the size of replacementdolomite rhombohedra (Zempolich, 1995). Microscaledolomite textures, which record the initial step-by-stepreplacement of limestone by dolomite and the neomor-phism and recrystallization of initial replacementdolomite fabrics, are distributed across dolomite–limestone transitions. These petrographic data definethe mechanism by which limestone was progressivelyreplaced by dolomite, and by which initial replacementfabrics were progressively recrystallized.

Replacement Dolomite

Initial replacement fabrics are found toward theperiphery of dolomite reaction fronts within partiallydolomitized limestone (Figure 16). Initial replacementdolomite is composed of calcian dolomite that con-tains inclusions of relic calcite (Zempolich, 1995). Twotypes of initial replacement styles are exhibited by theVajont dolomite: intragranular replacement—initialdolomitization begins with the selective dolomitiza-tion of ooids and other grains within oolite whereintergranular calcite cement has completely occludedpore space; and intergranular replacement—initialdolomitization begins with the dolomitization of ooidsand intergranular matrix (i.e., carbonate mud) prefer-entially along grain peripheries (Zempolich, 1995).Field and petrographic data suggest that these differ-ent replacement styles are dependent on the degree ofcementation within the precursor limestone fabric andoriginal carbonate mud content. Recrystallizedreplacement fabrics are found in completely dolomi-tized limestone nearest to faults and fractures.

Cathodoluminescent petrography and microprobeanalysis of initial replacement fabrics indicate thatreplacement and recrystallized dolomite found in boththe northern and southern study areas luminesces ahomogeneous dull red color, and that individual crys-tals are not compositionally zoned. Cathodolumines-cence also reveals that replacement dolomite crosscutsooid grains, pore-lining nonluminescent equant calcitecement, and banded-luminescent equant calcite spar

cement (Zempolich, 1995). The widespread uniformityin luminescence and lack of compositional zoning sug-gests that initial replacement by calcite-inclusion–rich,nonstoichiometric dolomite, and neomorphism ofthese phases to more stoichiometric compositions, wasthe product of one progressive dolomitization event(Zempolich, 1995). Postdolomitization processesinclude the recrystallization, dedolomitization, anddissolution of initial calcite-inclusion–rich replacementfabrics.

These petrographic observations are important forseveral reasons. First, a replacement origin for dolomitein the Vajont limestone is inferred by the pervasiveretainment of ooid ghosts in both dolomitized matrixand lithoclasts (Figure 16). These observations indicatethat precursor limestone was not wholly dissolved, andlater reprecipitated as dolomite in voids. Second, a latereplacement origin for the dolomite is indicated bycathodoluminescent study that indicates replacementdolomitization occurred sometime after calcite cemen-tation in burial settings.

Baroque Dolomite Cement

Baroque dolomite cement is found in associationwith replacement dolomite along both large-scale andsmall-scale fractures within the Mt. Grappa–Visentinanticline, and as massive pore fill within brecciatedcores of dolomite plumes located in the northern studyarea (Figures 11, 12). Baroque dolomite cement was notobserved within limestone or along fractures withinundolomitized limestone. This suggests that faults andfractures were the conduits by which dolomitizingfluid circulated, and that baroque dolomite cement wasa final pore-filling phase that precipitated after replace-ment dolomitization.

Regional Stable Isotopic Geochemistry

Compositions of replacement dolomite exhibit awide range of δ18O and a relatively narrow range of δ13Cvalues that overlap the Middle Jurassic marine carbon-ate compositions (Zempolich, 1995). Regional δ13C andδ18O compositions of replacement dolomite andbaroque dolomite cement are summarized in Figure 17.The 18O of replacement dolomite in northern dolomitelocalities is depleted relative to replacement dolomitelocated along the Mt. Grappa– Visentin anticline. Com-positions of baroque dolomite cement exhibit depleted18O compositions relative to replacement dolomite andMiddle Jurassic marine carbonate, and possess variablecarbon compositions. The 18O of baroque dolomitecement appears to be uniformly depleted throughoutthe region. These data, in addition to fluid inclusiondata (Th = 80–132°C, mean = 125°C, n = 12), suggestthat replacement dolomitization and baroque dolomitecementation occurred at elevated temperature, andthat dolomite replacement in the northern study areatook place at higher temperatures than that of thesouthern study area (Zempolich and Hardie, 1991a, b;Zempolich, 1995).

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 147

Page 161: Reservoir Quality Prediction in Sand and Carbonates

148 Zempolich and Hardie

Trace Elements

A characteristic of Vajont replacement dolomite inboth the southern and northern study areas is a low con-centration of Sr, Fe, and Mn (e.g., Col Visentin: Sr = 62.6ppm, Fe = 93.2 ppm, and Mn = 27.5 ppm; Vajont Dam: Sr= 32.1 ppm, Fe = 92.1 ppm, and Mn = 42.6 ppm) (Zem-polich, 1995). These values are similar to or are muchlower than estimates of “marine” dolomite (Sr = 50–850ppm, Fe = 10–2000 ppm, Mn = 5–275 ppm) (Al-Aasmand Veizer, 1982; Major, 1984; Saller, 1984; Aissoui, 1988;Dawans and Swart, 1988; Vahrenkamp and Swart,1990); “deep marine” dolomite (Fe = 2100 ppm, Mn =590 ppm) (Lumsden, 1988); late-stage recrystallized andburial dolomites (Sr = 35–147 ppm, Fe = 287–5115 ppm,Mn = 0.1–1069 ppm) (Montañez and Read, 1992; Mon-tañez, 1994); dolomites of various depositional settings(average Fe = 2790 ppm, Mn = 245 ppm) (Weber, 1964);and dolomites of hydrothermal brine origin (Gregg,1985; Gregg and Shelton, 1989). This comparison sug-gests that Vajont trace element compositions are notcompatible with dolomite replacement, neomorphism,or recrystallization involving fluids enriched in Sr, Fe,and Mn (i.e., burial fluids or hydrothermal brine). Whilerecrystallization of replacement dolomite and loss of Sr,Fe, and Mn through time is a possibility (Kupecz et al.,1993), most Vajont replacement dolomite exhibits petro-graphic evidence of initial replacement crystal frontsand engulfment of dissolution-resistant precursor calcite(Zempolich, 1995). The retention of these microfabricssuggests that geochemical compositions of replacementdolomite were emplaced during initial dolomitizationand not through recrystallization (Zempolich, 1995).

However, modeling of isotopic, trace element, andfluid inclusion data collected from both limestone anddolomite components indicate that Vajont stable iso-topic compositions and trace element concentrations arecompatible with initial dolomite replacement and neo-morphism or recrystallization by seawater-derived fluidat elevated temperature (Zempolich, 1995). If correct,these data and models may suggest that circulation ofseawater at temperatures ≤100°C may have causeddolomitization along faults and fractures within the Mt.Grappa–Visentin anticline (southern study area), andthat circulation of seawater and/or modified seawater attemperatures ≤200°C may have caused dolomitizationalong faults and fractures within synclines in the north-ern study area. Moreover, 87Sr/86Sr values of replace-ment dolomite from southern (87Sr/86Sr = 0.707104–0.707570; N = 9) and northern study localities (87Sr/86Sr= 0.707040–0.708180; N = 2) overlap model ranges thatutilize early Eocene and late Oligocene to early Mioceneseawater values (Zempolich, 1995). Collectively, theseresults suggest that (1) dolomitization of the Mt.Grappa–Visentin anticline occurred by the circulation ofEarly Tertiary seawater at temperatures of 35–100°Cconcomitant with initial early Eocene compression and(2) dolomitization of the northern dolomite localitiesoccurred by the circulation of Early to Middle Tertiaryseawater or modified seawater at temperatures ≤200°Cconcomitant with initial early Eocene or late Oligoceneto early middle Miocene compression.

Figure 16. Dolomite replacement textures (plane lightand cross-polarized light photomicrographs. (A, B)Partly replaced oolitic limestone. Replacement rhom-bohedra have preferentially nucleated within inter-granular matrix and along the periphery of ooids.Precursor ooid structures are defined by a greaterdensity of calcite inclusions within replacementrhombohedra (arrows). The dolomite–limestone con-tact between radial ooid cortices and replacementrhombohedra is sharp. (C) Replacement dolomite(partially recrystallized) with moldic pores “P”.Complete replacement dolomitization of ooliteresults in formation of ≤10%–15% porosity.

Page 162: Reservoir Quality Prediction in Sand and Carbonates

EVOLUTION OF POROSITY ANDRESERVOIR QUALITY THROUGH

DOLOMITIZATION

The complete dolomitization of Vajont oolite resultsin the replacement of calcite ooids and the formation ofpartly oomoldic and intercrystalline pores (e.g., Figures16, 18). Partial oomoldic fabric forms through the com-plete replacement of ooid grains by medium to coarselycrystalline dolomite. Intercrystalline porosity developsas a result of the replacement of fine oolite and mudmatrix. Visual estimates of porosity within dolomitizedoolite range to 10%–15% in thin section. This is in agree-ment with a theoretical 13% increase in porosity throughthe volume-for-volume replacement of calcite bydolomite (Weyl, 1960). The distribution of moldic poreswithin grain interiors and the development of someintercrystalline porosity suggest that the pores in initial

replacement dolomite fabrics were relatively isolated,and that the permeability developed at this stage ofdolomitization was relatively low (inferred permeabili-ties of ~1–100 md; Lucia, 1995).

The macroscale rearrangement of moldic pores toform separate-vug and touching-vug pore space (Lucia,1995) is first observed as a progression from moldicpores to separate vugs (nonfabric-selective pores) incompletely replaced grainstone and packstone (Figure18). The retainment of ooid ghosts around the marginsof separate vugs indicates that enlargement of moldicpores occurred through local pore migration and crystalrearrangement (Zempolich, 1995). Next, a transitionfrom separate-vug fabric to touching-vug fabric indi-cates that continued dissolution and recrystallizationcaused separate vugs and crystalline material tomigrate and align, thereby forming alternations oftouching-vug and dense recrystallized fabrics (e.g., Fig-ure 18E, F). The development of interconnected pores in

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 149

Figure 17. Regional distribution of Vajont dolomite and oxygen isotopic compositions. Compositions of replace-ment dolomite in the northern study area are more depleted in 18O relative to those of replacement dolomite in thesouthern study area. Compositions of baroque dolomite cement in the northern study area are slightly depleted in18O relative to those of baroque dolomite cement in the southern study area (Zempolich, 1995). These and othergeochemical and petrographic data suggest that dolomitizing fluids in the north circulated at higher temperatures(75–175°C) than did those in the south (35–100°C). B = baroque dolomite cement; R = replacement dolomite.

Page 163: Reservoir Quality Prediction in Sand and Carbonates

150 Zempolich and Hardie

touching-vug fabric suggests that permeability hasbeen increased without a change in the total porosity ofthe matrix (inferred permeabilities ≥100 md; Lucia,1995). The retainment of ooid ghosts in these fabrics isless common and indicates a progressive-phase separa-tion between pores and crystalline material. This is acharacteristic of the sintering process (Barrett, 1987;Zempolich, 1995).

The significance of progressive textural modifica-tion in Vajont dolomite is that while reservoir-gradeporosity may be formed through the initial replace-ment of limestone by dolomite (≤10%–15%), reser-voir-grade permeability is created through therecrystallization of intial replacement dolomite andpores. Given the rock volumes of the large-scaledolomite wedge and dolomite plumes (i.e., 50–94km3 and ≥2 ×10–2 km3, respectively) found in the pres-ent study area, such pore space and enhanced perme-ability could potentially form significant economichydrocarbon accumulations. For example, assuming10% porosity, the large-scale dolomite wedge (Mt.Grappa–Visentin anticline) may contain ≤3 to 6 bil-lion bbls of pore space, while the typical dolomiteplume (e.g., Vajont Canyon) may contain ≥ 12.5 mil-lion bbls of pore space. Importantly, the large size ofVajont dolomite bodies and the formation and redis-tribution of porosity through late-stage replacementdolomitzation and recrystallization illustrate that sig-nificant dolomite reservoirs may be created throughmassive, late-stage, fault-related burial dolomitiza-tion.

THE ORIGIN OF THE VAJONT DOLOMITE

Field, petrographic, and geochemical data point todolomitization of the Vajont Limestone by regional-scale circulation of Tertiary seawater within anticlinesand synclines that were formed during TertiaryAlpine deformation. Theoretical and laboratory circu-lation models and fluid flow patterns observed inmodern thrust zones are consistent with this interpre-tation, as discussed below.

Physiochemical Factors

As reviewed by Hardie (1987), a number of physio-chemical factors influence the formation of dolomite insedimentary and burial environments. These factorsinclude thermodynamics, kinetics, mass transfer, andthe nature of the precursor host rock. Current modelsof dolomitization, such as mixing-zone, tidal flat,evaporative-brine, and schizohaline, have inherentweaknesses with regard to one or more of these fac-tors, the more serious of which are related to thermo-dynamics, kinetics, and the mass transfer of Mg. Theseproblems are easily overcome at elevated temperatureand within flow regimes capable of circulating largeamounts of Mg-bearing fluid (Hardie, 1987; Wilson,1989; Wilson et al., 1990).

With regard to the Vajont dolomitization, we candraw the following conclusions about the physio-chemical factors involved:

1. The widespread occurrence of reaction frontsbetween dolomite and precursor limestone at alllocalities indicates that dolomitizing fluids wereoversaturated with respect to dolomite and under-saturated with respect to calcite. Initial replacivedolomitization of limestone, therefore, most prob-ably occurred through the general reaction:

2 CaCO3(cal) + Mg2+(aq) <—-> CaMg(CO3)2(dol) + Ca2+(aq) (1)

2. Field and petrographic data show that dolomitefronts moved out and away from fluid conduits(i.e., faults and fractures).

3. The 18O compositions of replacement dolomite insouthern localities are relatively enriched com-pared with those of northern dolomite localities,and suggest that dolomitization occurred atmoderate temperatures (≤100°C).

4. The 18O compositions of replacement dolomite innorthern localities are depleted relative to MiddleJurassic marine carbonate and, together with fluidinclusion data, suggest that dolomitizationoccurred at more elevated temperatures (≥125°C).

These factors indicate that the dolomitizing fluidswere introduced to the Vajont basinal limestones alongfractures and faults at elevated temperatures, and thatdiffusion of Mg through relatively nonporous lime-stone resulted in the formation of massive replacementdolomite (Zempolich and Hardie, 1991a, b; Zempolich,1995).

Flow Volumes and Delivery of Magnesium

To get some measure of the mass transfer require-ments, the volumes of fluid necessary for the dolomiti-zation of the large wedge within the Mt. Grappa–Visentin anticline (50–94 km3 dolomite) was calculated.With seawater as the dolomitizing fluid, the calculationyields 2.74 ×104 km3 at 35°C and 2.53 ×105 km3 at 100°C(Zempolich, 1995). Such large volumes demand thatdynamic transport of Mg from an external source musthave occurred. For scale, the fluid volume calculated forthe seawater case at 100°C and a dolomite rock volumeof 50 km3 is equal to the volume of a small sea, 1 kmdeep and 165 ×165 km in area. Although smaller, the cal-culated volumes of the fluid necessary for the formationof a representative dolomite plume (2.36 ×10–2 km3

dolomite) found in the northern study area (e.g., seawa-ter test case at 200–300°C = 6.6–12.9 km3 seawater) (Zem-polich, 1995) are impressive, and demand the dynamicflow of Mg-bearing fluid to promote dolomitization inthe northern study area.

What remains to be explained is what kind of Mg-bearing fluid was involved, where massive quantitiesof this fluid were generated, and how the fluid wastransported to the network of large- and small-scalefluid conduits.

Page 164: Reservoir Quality Prediction in Sand and Carbonates

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 151

Figure 18. Dolomite textures and associated pores formed through the progressive replacement of limestoneand the progressive recrystallization of intermediate replacement fabrics (Villanova section). (A, B) Close-upphotographs of the left side of the reaction front that is depicted in Figure 15 (note hammer). The reactionfront is separated into discrete zones of dolomite textures and pores oriented subparallel to the fault.Starting from the fault marked “F” these include: (1) densely crystalline dolomite with minor amounts ofpore space; (2) recrystallized dolomite with touching-vug pores; (3) recrystallized and neomorphic dolomitewith separate-vug and moldic pores (Figure 16C); (4) initial replacement dolomite in partially dolomitizedlimestone. (C) Close-up field photograph of moldic and separate-vug fabric (3). (D) Polished-slab photo-graph of moldic and separate-vug fabric (3). Visual estimates of porosity in both slab and thin section rangeto 10%–15%, inferred permeabilities range from 1–100 md (e.g., Lucia, 1995). (E) Close-up field photograph oftouching-vug fabric (2). Bands of touching-vug pores are located ~1–2 cm apart and are oriented parallel tothe reaction front. (F) Polished-slab photograph of touching-vug fabric (2) (sample oriented so that reactionfront is to the left). Vugs are ≤1 cm in length and are laterally interconnected. Visual estimates of porosity inboth slab and thin section range to 10%–15% and are similar to those estimated for moldic and separate-vugfabric. Inferred permeabilities are ≥100 md (Lucia, 1995).

Page 165: Reservoir Quality Prediction in Sand and Carbonates

152 Zempolich and Hardie

Models for Dolomitization of the Vajont Limestone

Dolomite distribution in the Belluno Basin is con-fined to Mesozoic slope and basinal sediments. Thedolomitization of such a thick sequence of relativelyimpermeable lithologies is unusual in that primarydepositional porosity and permeability were negligi-ble. This presents a major problem for the transporta-tion of large quantities of Mg-bearing fluid to the sitesof reaction because primary large-scale fluid conduits,such as permeable siliciclastic sands or carbonates, areabsent in these deep-water sediments.

A number of hydrologic models have been devel-oped or invoked to explain the occurrence of dolomitein shallow platform settings. These models have beenbased largely on the occurrence of dolomite inHolocene environments and have been applied toancient dolomites by analogy with facies and paleo-geographic settings. Shallow dolomitization modelsinclude mixing zone (Hanshaw et al., 1971; Badioza-mani, 1973; Hanshaw and Back, 1980; Sandford, 1987),reflux (Adams and Rhodes, 1960; Sears and Lucia,1980; Simms, 1984; Whitaker et al., 1994), tidal pump-ing (Carballo et al., 1987), and evaporative pumping(McKenzie, 1981; Ruppel and Cander, 1988). Thesehydrologic and depositional models predict the occur-rence of dolomite in shallow shelf and platform mar-gin settings (Kaufman, 1994), and so cannot explainthe dolomitization of the Vajont Limestone, which wasdeposited as a thick sequence of carbonate gravityflows in slope and basin settings of the Belluno Basin.

Other hydrologic models and settings for dolomiti-zation include contemporaneous dolomitization ofdeep marine sediments by cool ocean water (Baker andBurns, 1985; Lumsden, 1985, 1988; Mullins et al., 1985),topographic-driven flow (Garven and Freeze, 1984;Garven, 1985; Gregg, 1985; Barrett, 1987; Ge and Gar-ven, 1989; Yao and Demicco, 1995), compaction-drivenflow (Jodry, 1969; Mattes and Mountjoy, 1980; Bethke,1985), and thermally driven flow (Elder, 1965; Kohoutet al., 1977; Simms, 1984; Aharon et al., 1987; Wilson etal., 1990; Kaufman, 1994).

Deep-marine sedimentary dolomitization is anunlikely explanation for massive dolomitization of theVajont limestone because (1) dolomite bodies crosscutbasinal stratigraphy, (2) Vajont replacement dolomitedisplays relatively coarse textures and depleted oxy-gen compositions that are quite different from the fine-grained disseminated dolomite that characterizesthese occurrences (Lumsden, 1988), and (3) deep-water dolomite is a volumetrically minor component(average 0.5%) of modern deep-water sedimentarycover (Lumsden, 1988).

Topographic-driven flow may result in the long-term or transient flow of fluid in basins through thedevelopment of sufficient recharge and hydrostatichead in neighboring uplift areas (Garven and Freeze,1984; Garven, 1985; Ge and Garven, 1989). However,such a model is unlikely for dolomitization of theVajont limestone because: (1) uplift of the VenetianAlps and the formation of a possible recharge areaduring the middle Miocene (Massari et al., 1986;

Doglioni, 1990) included the basinal sedimentary sec-tion that is now dolomitized (i.e., Mt. Grappa–Visentinanticline, Belluno thrust); (2) initial thrust movementoccurred synchronous with deposition of early Eoceneto middle Miocene marine siliciclastics, marls, and car-bonates in the Venetian Basin (as far north as Cortina);(3) dolomite bodies are aligned parallel to and arehosted within structural axes formed during initialcompression of the Venetian Alps and prior to signifi-cant uplift; and (4) basin-scale topographic flow ema-nating from the Appenine Mountains to the south andmigration through the Po Basin into the study areawould have produced isotopic trends opposite tothose observed (i.e., depleted 18O values in the south,enriched 18O values in the north).

The compaction and dewatering of shales is anunlikely source of dolomitizing fluids because shale isa volumetrically minor component of Late Paleozoicand Mesozoic sediments of the area, and because ofthe enormous volume of Mg-bearing fluid, whichmust be accounted for by mass balance calculations.Kohout convection (Simms, 1984) would predict theoccurrence of dolomite in platform margin to periplat-form settings, which is a dolomitization pattern that isunsupported by field, sedimentologic, and strati-graphic evidence.

Considering the discussion above and the collectivefield, petrographic, and geochemical evidence thatsuggests that dolomitization of the Vajont Limestoneoccurred by the large-scale circulation of fluid at ele-vated temperature along faults and fractures, dolomi-tization most likely occurred through the circulation ofhydrothermal fluids at depth. Considering the low Sr,Fe, and Mn concentrations in replacement dolomiteand a lack of associated Mississippi Valley-type miner-alization, dolomitizing fluids must also have been lowin Sr, Fe, Mn, and other metals, yet were capable oftransporting large quantities of Mg. Geochemical dataand modeling (see below) suggest that hydrothermaldolomitization most likely occurred through the circu-lation of seawater or modified seawater at depth(Zempolich, 1995).

Thermal Convection

As argued above, the fluid volumes that arerequired to produce both the extensive dolomite bodyin the southern study area and the narrow isolateddolomite plumes in the northern study area requirethat large-scale fluid transport must have occurred.Given that the fluid inclusion and geochemical dataindicate that dolomitization of the Vajont limestoneoccurred at elevated temperature, it is likely that mas-sive volumes of Mg-bearing fluid were delivered to thedolomitization sites by thermal convection, and thatthe geometry of these convective cells was dependenton temperature and availability of fluid conduits.

Through study of dolomitization patterns in theLatemar buildup, an isolated Late Triassic carbonateplatform penetrated by rift-related Late Triassic volcanics, Wilson et al. (1990) have proposed severalthermal-convective flow models to explain the

Page 166: Reservoir Quality Prediction in Sand and Carbonates

occurrence of a massive mushroom-shaped dolomitebody 1–2 km2 in diameter. Critical to the generation ofconvective flow in these models is the presence of a heatsource and a supply of Mg. Using other field and geo-chemical data, Wilson et al. (1990) conclude thatdolomitization was most likely caused by the thermalconvection of Late Triassic seawater driven by local ele-vated temperatures due to a volcanic intrusion. Convec-tive models similar to the one proposed for the Latemarbuildup (e.g., Kaufman, 1994) are unsuitable explana-tions for dolomitization of the Vajont limestone because(1) volcanic intrusions are not known in either thesouthern or northern study areas, (2) dolomite bodiesare distributed within basinal rather than platformstrata, (3) dolomite bodies are associated with compres-sional structures, and (4) multiple dolomite plumes arefound along the same structural trend. Thus, if theVajont dolomite geometries were produced by the ther-mal convection of Mg-bearing fluid, other convectionmodels must be called upon to explain the unusualdolomite geometries that are now found in basinalstrata contained within compressional structures.

Vajont Dolomite Geometries andTheoretical Convective Flow Patterns

Convection models that may be applicable todolomite geometries observed in the Vajont and otherMesozoic basinal sediments have been explored byElder (1965, 1967, 1977). Elder’s models were devel-oped through two-dimensional numerical simulationsand scaled laboratory experiments to simulate the cir-culation and mass discharge of fluid in geothermaland volcanic areas, rifts, and oceanic rises (Figure 19).These scaled models rely on linear, basal heat sources,and approximate the physical dimensions of zones offracture-enhanced permeability that are found in thefaulted crests and troughs of anticlines and synclinesthat are host to Vajont dolomite bodies (Zempolich,1995). The theoretical convective flow patterns of thesemodels predict the occurrence of multiple isolatedplumes and large-scale flow geometries that approxi-mate the dolomite geometries that are observed in thenorthern and southern study areas (Figure 20).

Isotopic, trace element, and fluid inclusion evidencesuggests that dolomite bodies in the northern andsouthern areas evolved under different thermal regimes(Zempolich, 1995). As dolomite bodies in the south andnorth are hosted in a similar succession and thickness ofbasinal strata, it is unlikely that this temperature differ-ence arose from differences in burial history. Tempera-ture differences due to an extraneous heat source, suchas the intrusion of magma, can be ruled out becausesuch intrusions are not present in the study areas. Thus,the temperature differences must have resulted fromsome other thermotectonic perturbation.

The difference in temperature regimes may berelated to the depth that structural deformationreached and to patterns of fluid circulation. For exam-ple, in the southern study area, balanced reconstruc-tions of Doglioni (1990) suggest that the Mt.Grappa–Visentin anticline detached along a shallow

decollement (≤5 km deep) within the Mesozoic section,whereas thrusts in the northern study area detachedalong fairly deep decollements (5–10 km) within theMesozoic section and Paleozoic basement. Initialdetachment and thrusting took place prior to the earlymiddle Miocene during development of the VenetianAlps foredeep (Massari et al., 1986; Doglioni, 1990).Within this subaqueous thrust system, anticlines andsynclines were dissected by numerous vertical-subver-tical faults (Figures 9, 10). Assuming a normal thermalgradient of ~30°C/km, ambient temperatures at thedepth of the decollement horizons (5 and 10 km) wouldbe ~150° and 300°C. These observations suggest thatfluids, circulating downward along faults and fractureswithin anticlines and synclines, may have been heatedby conduction to temperatures approaching 150°C and300°C, respectively. These postulated temperature dif-ferences are consistent with geochemical data and achange in the geometry of the dolomite bodies fromnarrow isolated plumes in the north to a broaddolomite wedge in the south (Zempolich, 1995). There-fore, it is postulated that vertical to subvertical faultswithin anticlines and synclines produced hydrologicconduits that connected overlying Tertiary seawaterwith deeply buried Mesozoic basinal sediments,thereby creating large-scale convective hydrologic sys-tems, which enabled the massive dolomitization of oth-erwise tight basinal sequences (Figure 20).

In summary, massive dolomitization of the VajontLimestone by the convective circulation of Early toMiddle Tertiary seawater is suggested by a consis-tency among dolomite and structural field relation-ships, petrographic data, geochemical data, andtheoretical hydrologic models. It is proposed that thedelivery of Mg-bearing fluid and massive replacementdolomitization was promoted by a combination of: (1)large-scale fluid flow along Tertiary compressionalstructures that provided the main plumbing by whichan Mg-rich reservoir (seawater) was put in communica-tion with a deep heat source. The ensuing generation ofthermal-convection cells within anticlines and synclinesultimately controlled the overall shape and distributionof large-scale dolomite bodies; and (2) small-scale fluidflow emanating from large-scale flow systems. Small-scale fluid flow along faults, fractures, and beddingplanes controlled the propagation and orientation ofreaction fronts. In this manner, vast quantities of Mgwere delivered to relatively impermeable basinal sedi-ments of the Belluno Basin at elevated temperature andat a broad range of scales. Kinetic barriers involved inthe formation of dolomite were overcome by the ele-vated temperature and the high Mg/Ca ratio of seawa-ter (Hardie, 1987).

MODERN HYDROLOGIC ANDSTRUCTURAL ANALOGS

The proposed model for massive dolomitization ofthe Vajont Limestone depends on the thrusting of thicksequences of limestone in a marine environment, the

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 153

Page 167: Reservoir Quality Prediction in Sand and Carbonates

154 Zempolich and Hardie

development of a deeply rooted fault and fracture net-work through the crests and troughs of thrust-relatedanticlines and synclines, and the generation of large-scale convective fluid flow and diffusion of Mg throughlimestone matrix. It is proposed that large-scale fluidflow systems developed along fracture sets located

within anticlinal and synclinal structures, and that thesefracture networks controlled the overall geometry anddistribution of dolomite bodies. Many of these struc-tural and hydrologic features are present in modernsubduction or transpressive compressional zoneswhere carbonate and siliciclastic sediments are

Figure 19. Theoretical convective fluid flow models for the circulation of heated fluid in two-dimensionalporous mediums (modified from Elder, 1977). 1: Fluid flow model with restricted upward outflow. The ther-mal interface is uniformly heated from below, and surface discharge is localized within the middle of theupper surface. (A) Theoretical isotherms. (B) Hypothetical dolomite geometries resulting from convectivefluid flow (A), assuming that the kinetic inhibitions of dolomitization are overcome at elevated temperature.It is postulated that similar convective flow of Mg-bearing fluid, first downward and then upward along frac-tures within the axes of synclines, may have given rise to the multiple occurrence of dolomite plumes that arenow present in the northern study area (compare with Figures 3, 8, 11–13, and 20). 2: Fluid flow model withenhanced lateral outflow (to the right). The thermal interface is uniformly heated from below. Regional fluidflow is from left to right. (A) Theoretical isotherms. (B) Hypothetical dolomite geometry resulting from con-vective fluid flow (B), assuming that the kinetic inhibitions of dolomitization are overcome at elevated tem-perature. It is postulated that similar large-scale convective flow of Mg-bearing fluid may have given rise tothe large, thickening-westward dolomite wedge that is now present along the Mt. Grappa–Visentin anticlinein the southern study area (compare with Figures 3, 8–10, and 20).

Page 168: Reservoir Quality Prediction in Sand and Carbonates

deformed into a series of thrusted anticlines and syn-clines. For example, shallow- and deep-water carbonateof Mesozoic and Tertiary age is found in an extensivethrust zone located along the southeast Bahamas–His-paniola collision zone (Ditty et al., 1977; Austin, 1983;Mullins et al., 1992). This submerged thrust zone iscomposed of a series of anticlines and synclines at sub-sea depths of ~1000–3000 m. Interestingly, gas-hydratezones and “groundwater seeps” identified by high seis-mic data reflectivities (Austin, 1983; Mullins et al., 1992)are present in some, but not all, of the crests and hang-ing walls of anticlines (Figure 21). These features aredistributed in the same structural position as the large-

scale dolomite front within the Mt. Grappa–Visentinanticline. Moreover, crestal positions of these anticlinesare extensively dissected by large-scale vertical to sub-vertical “keystone” faults (Figure 22) (Austin, 1983).The buried geometry of these anticlines and their asso-ciated faults are remarkably similar to the structural set-ting and fault pattern recognized in the Mt.Grappa–Visentin anticline. Moreover, the subsea sedi-mentologic setting of the southeast Bahamas–Hispan-iola collision zone is similar to the Tertiary foredeepformed during initial thrusting of the Venetian Alps.These data, together with previously presented field,petrographic, and geochemical data, support the idea

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 155

Venetian Basin

N

Thermal convectioncells oriented alongfaults and fractures

25 kilometers

Large-ScaleDolomite

Front

IsolatedDolomitePlumes

THEORETICAL FLUID FLOW PATTERNSDURING DOLOMITIZATION OF THE

VAJONT LIMESTONETERTIARY SUBAQUEOUS

THRUST ZONES

TertiarySeawater

Monte G

rappa-

Visentin

Antic

line

DownwardInfiltratingSeawater Alpine Fore

deep

Figure 20. Proposed flow pathways of dolomitizing fluids and their relation to structural features of theVenetian Alps thrust belt (fluid flow patterns after Elder, 1977; present-day structural configuration afterDoglioni, 1990). It is proposed that massive dolomitization of the Vajont Limestone and other Mesozoic basi-nal sequences was a consequence of the thermal convection of Early to Middle Tertiary seawater along faultsand fractures that were formed during initial tectonic deformation and that breached the Tertiary sea floor.Structural and crosscutting stratigraphic relationships of late-stage dolomite suggest that dolomitizationoccurred sometime during the early Eocene/late Oligocene or late Oligocene/early middle Miocene, and priorto significant uplift of the basinal succession that occurred during the early middle Miocene to Pliocene.Convective fluid circulation is postulated to have developed as a result of extensive fracturing and faultingwithin the axes of synclines and anticlines and the downward infiltration of seawater. At depth, the seawaterwas presumably heated by conductive heat flow and then driven upward along the fracture and fault networkby buoyant forces (compare with Figures 3, 17, 21, and 22).

Page 169: Reservoir Quality Prediction in Sand and Carbonates

156 Zempolich and Hardie

that thrusting of the Venetian Alps during the Tertiaryand the concomitant thermal circulation of seawaterthrough faults and fractures within the rising fold beltwas responsible for the occurrence of bodies of massivedolomite within the Vajont Limestone.

Recent study has identified that fluid expulsion andventing along vertical strike-slip faults and crestalfaults in thrust-related anticlines of the Cascadia accre-tionary prism is significant (Kulm et al., 1986; Ritger etal., 1987; Carson et al., 1990; Sample et al., 1993; Tobinet al., 1993). Flow rates of 100 m/yr have beendescribed for these systems, suggesting that large-scale flow has developed in response to the dewater-ing of prism sediments and expulsion of fluid throughvertical channelized flow. Calcite cements sampledfrom siliciclastic sediments outcropping along thesefault zones possess depleted oxygen compositions,enriched to depleted carbon compositions, and radi-ogenic Sr compositions that are thought to representprecipitation at temperatures as great as 100°C (calcitecement δ18O = –4 to –13‰, δ13C = –1 to –25‰, 87Sr/86Sr= 0.70975–0.71279) (Sample et al., 1993).

The origin and composition of these cements hasbeen attributed to the complex interaction of deeplyderived interstitial pore fluid with clays, thermogenicmethane, and marine water (Sample et al., 1993).Despite the marine influence observed in these cementsand the distribution of cements in the crests of anticlines,the origin of these fluids has been solely attributed to thedewatering of the accretionary complex. An alternativeexplanation for the occurrence of channelized fluid flowwithin these structures would be the thermal-driven cir-culation of seawater through these extensive fault andfracture systems. In such a scenario, the modification ofsome seawater would presumably occur through reac-tion with siliciclastic sediment and organics. If correct,channelized fluid flow along faults and fractures, andthe chemical modification of seawater in the Cascadiaaccretionary prism, would be analogous to that pro-posed for the dolomitization of the Vajont limestone,including the relationship of dolomite bodies to struc-ture and large-scale fluid flow and marine to nonmarinecompositions of replacement dolomite.

The similarity between the Mt. Grappa–Visentinanticline and Belluno thrust with these modern struc-tures supports the proposed hydrostratigraphic modelthat involves the convection of seawater along linearzones of high permeability. The southeast Bahamas–Hispaniola collision zone and Cascadia accretionaryprism may be modern structural and hydrostrati-graphic analogs for the tectonic deformation and thedolomitization of the Vajont and other Mesozoic basinalsediments during the Tertiary.

POTENTIAL DOLOMITE RESERVOIRANALOGIES

Field mapping, petrography, and geochemistry ofthe Vajont dolomite reveal a strong relationshipbetween hydrothermal dolomitization and tectonism.The enhanced porosity and permeability within thesedolomite bodies suggests that these bodies may wellrepresent exhumed dolomite reservoirs created in tec-tonically deformed carbonate provinces. Dolomitereservoir geometries illustrated in this study include:meter-scale dolomite bodies located parallel to fracturenetworks; multiple isolated dolomite plumes, severalhundred meters in width and height, located alongstructural trends; and large-scale dolomite bodies, kilo-meters to tens of kilometers in scale, encompassing thecrests of anticlines. Accordingly, the recognition ofthese dolomite geometries and this style of dolomitiza-tion in subsurface settings may define new explorationtargets in the search for oil and gas, and/or provide ana-log geometries for reservoir characterization (Figure 23).

Subsurface examples of ellipsoidal, areally restricteddolomite bodies apparently associated with tectoniclineaments are postulated to exist in several basins. Forexample, in Paleozoic strata of the Michigan Basin,small-scale ellipsoidal dolomite reservoirs are alignedNW-SE in association with the dominant NW-SE fracture network imposed on the basin duringAppalachian orogenesis (Prouty, 1983). Dolomite iso-pleths suggest that these bodies thin away rapidly from

Figure 21. Subaqueous thrust belt (~1000–3000 msubsea) north of Hispaniola (modified from Mullinset al., 1992). Shallow- and deep-water carbonates ofMesozoic and Tertiary age are thrusted into a seriesof anticlines and synclines due to regional transpres-sion. “Groundwater seeps” (circles) and gas-hydratezones, which are identified by high seismic datareflectivities, are present along and parallel to someof the crests and hanging walls of anticlinal thrusts.These structural and hydrostratigraphic settings arethought to be analogous to those that promoted thedolomitization of the Vajont Limestone and otherMesozoic basinal sediments during deformation ofthe Venetian Alps during the Tertiary.

Page 170: Reservoir Quality Prediction in Sand and Carbonates

tectonic lineaments. Dolomite types include fine-crys-talline dolomite and late coarse-crystalline and baroquetextures (Zempolich, 1984, personal observation).Plume-shaped isolated dolomite reservoirs are alsofound in Paleozoic rocks in the Williston Basin (R.D.Perkins, 1991, personal communication). By analogywith dolomite bodies observed in the Vajont Limestone,it is possible that these dolomite bodies were producedby the convection of heated fluid(s) along deep-seatedtectonic fractures and faults. If correct, these basin andoutcrop examples would predict that significantdolomite reservoirs may be hosted in carbonate stratathat have been deformed in peri- and intracratonic tec-tonic settings and subjected to hydrothermal dolomiti-zation processes. Such hydrothermal dolomite bodiesmay be difficult to recognize in sequences that have alsobeen affected by early shallow dolomitization processes.

The delineation of hydrothermal dolomite bodies inthe subsurface may include (1) the mapping of dolomiti-zation fronts using dolomite abundances calculatedfrom well logs and core, and (2) the identification ofthermal diagenetic fluids and textural trends using pet-rographic and geochemical techniques. Once thehydrothermal process has been delineated, the searchfor new exploration targets may be concentrated on anti-clinal and synclinal structures within buried fold and

thrust belts, and along zones of deep-seated tectonicfractures and faults within intracratonic basins.Dolomite plumes may be identified by seismic reflectiondata methods due to the disruption of bedded limestoneby crosscutting dolomite fronts and by formation ofbreccia cores. The seismic reflection data expression ofsuch dolomite bodies would, theoretically, consist of a“rootless,” oriented chaotic zone (several hundredmeters high) interspersed within layered reflectors (i.e.,undolomitized bedded limestone).

SUMMARY AND CONCLUSIONS

Through field and laboratory study of Vajont lime-stone and dolomite, a number of inferences can bemade as to the formation of massive replacementdolomite and formation of dolomite reservoirsthrough late-stage fault-related, burial dolomitiza-tion. Field distribution of dolomite bodies and petro-graphic and geochemical data collectively suggestthat massive replacement dolomitization occurred asa result of the circulation of hot Mg-bearing fluidspiped into the Vajont and other Mesozoic basinal sed-iments along a master network of faults and fractures.The faulting and fracturing of Mesozoic basinal sedi-ment is related to Alpine thermotectonics, which

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 157

Figure 22. Schematic cross section of the southeast Bahama–Hispaniola collision zone (modified from Austin,1983). Subaqueous thrusts are composed of shallow- and deep-water carbonates of Mesozoic–Tertiary age andare found at present-day subsea depths of 1000–3000 m. Numerous vertical to subvertical faults dissect anti-clines and synclines that were formed during regional transpression. It is postulated here that extensive frac-turing and faulting within the axes of these submerged structures may allow for the downward infiltration ofseawater. Potential zones of dolomitization may exist in the cores of these anticlines and synclines (stippledpattern) due to the thermal-convective circulation of seawater upward and along these extensive fracture andfault networks (compare with Figures 3, 21).

Page 171: Reservoir Quality Prediction in Sand and Carbonates

158 Zempolich and Hardie

formed a series of thrust-related anticlines and syn-clines in the study area during early Eocene? and lateOligocene to Recent times. These faults and fracturesopened up porous and permeable pathways withinan otherwise tight sequence of basinal limestone.Dolomitization is postulated to have occurred con-comitant with initial thrusting during early Eoceneand/or late Oligocene to early middle Miocene timewhile the study area was still inundated by seawater.Dolomitization was completed prior to rapid upliftand subaerial exposure of the region during the middle Miocene to Pliocene. Rapid uplift followinginitial deformation and dolomitization preservedmetastable dolomite textures and compositionsacross limestone–dolomite transitions.

It is proposed that circulation of seawater was dri-ven by both large- and small-scale transport processesthat controlled the shape and distribution of dolomitebodies, reaction fronts, and replacement styles. Large-scale fluid movement involved the thermal convectionof Tertiary seawater through anticlines and synclines.In addition to Tertiary structures, dolomitizing fluidsalso utilized inherited structural elements such aspaleolineaments and paleosynsedimentary breccia.Convection cells were developed parallel to the axes ofthese structures through extensive subvertical to verti-cal faults and fractures. In the southern study area,fluid convection resulted in the formation of a large-scale dolomite body that is ~25 km long, 10 ×15 kmwide, and ≥400–500 m thick. In the northern study

Figure 23. Summary of Vajont dolomite bodies that are found in outcrop of the Venetian Alps. The diagram onthe left is a schematic cross view of these bodies; the diagram on the right is a schematic map view of thesebodies if projected into the subsurface. Vajont dolomite bodies are potential analogs for dolomite reservoirscreated in subsurface settings due to the formation of porous and permeable bodies in otherwise-tight deep-water limestone. Potential reservoir geometries that were created in the Vajont Limestone by massive replace-ment dolomitization and recrystallization in association with tectonism and the hydrothermal circulation ofMg-bearing fluid include: (A) small-scale dolomite wedges (meters to tens of meters in width) oriented paral-lel to subparallel with faults and fractures; (B) multiple isolated dolomite plumes (200–300 m wide, 300–400 mhigh) cored by dolomite breccia and located along structural trends; and (C) large-scale dolomite bodies (10–20km long, 5–10 km wide, and ≥400 m thick) located in the crests of major anticlines.

Page 172: Reservoir Quality Prediction in Sand and Carbonates

area, fluid convection resulted in the formation ofmultiple rootless dolomite plumes that are >300 mhigh and ~100–200 m wide.

Replacement of limestone by dolomite occurred bythe microscale dissolution of precursor limestone andprecipitation of dolomite. Complete dolomitizationformed porous moldic and intercrystalline fabrics withporosities up to 10% to 15%, and inferred permeabili-ties of 1–100 md. Recrystallization progressively amal-gamated moldic and intercrystalline pores anddolomite to form separate-vug, touching-vug, anddense crystalline fabrics with inferred permeabilities≥100 md. Consistency in the development of dolomitetextures, dolomite composition, and porosity withrespect to limestone–dolomite transitions suggests thatthe massive replacement of limestone by dolomite, andthe formation of reservoir-grade porosity and perme-ability, occurs through a predictable pattern of replace-ment and recrystallization.

Dolomite geometries in the northern and southernstudy areas are consistent with theoretical circulationmodels that predict the formation of large-scale flowsystems and the multiple occurrence of isolatedplumes due to the thermal convection of fluid. The pro-posed thermotectonic model for the formation of mas-sive replacement dolomite in the Vajont Limestonemay have modern analogs in active thrust zones of thesoutheast Bahamas and the Pacific Northwest.

The geometry, size, and distribution of dolomitebodies within the Vajont Limestone and other Meso-zoic basinal sediments indicate that late-stage thermo-tectonic dolomitization is an important process bywhich massive replacement dolomite may form.Moreover, these examples illustrate that both large-and small-scale dolomite reservoirs may be createdthrough late-stage dolomitization. Similar bodies inthe subsurface may prove to be attractive explorationtargets.

ACKNOWLEDGMENTS

Daniele Masetti, Carlo Doglioni, and AlfonsoBosellini of the University of Ferrara provided logisti-cal support that made this study possible. Dmitri Sver-jensky, Grant Garven, Owen Phillips, and Saki Olsen(Johns Hopkins University) provided help andinstruction on many of the geochemical and hydro-logic concepts that were evaluated during the courseof this study. K.C. Lohmann and Jim Burdett (Univer-sity of Michigan) provided carbon and oxygen isotopicanalyses, and Lynn Walters and Ted Huston (Univer-sity of Michigan) provided trace element (ICP) analy-ses. Tim Denison and Mobil Oil Corporation providedSr isotopic analyses. Special thanks are extended toAGIP for providing access to core from the Belluno 1well, and to ENEL for access to the Vajont Dam area.This study benefited from the support and help ofmany family members, fellow students, and friends,including Michele Claps, Paul A. Dunn, Linda A. Hin-nov, Joseph B. Paul, and Lyndon A. Yose. We would

like to thank J.A. Kupecz and J.R. Markello for provid-ing critical review of this manuscript. This study wasmade possible by grants from the American Associa-tion of Petroleum Geologists, the Geological Society ofAmerica, Sigma Xi, Mobil Oil Corporation, The JohnsHopkins University Balk Fund, and the National Sci-ence Foundation (Grant #EAR910510).

REFERENCES CITED

Adams, J.E., and M.L. Rhodes, 1960, Dolomitization by seepage refluxion: AAPG Bulletin, v. 44, p. 1912–1921.

Aharon, P., R.A. Socki, and L. Chan, 1987, Dolomitiza-tion of atolls by sea water convection flow: test of ahypothesis at Niue, South Pacific: Journal of Geology,v. 95, p. 187–203.

Aissaoui, D.M., 1988, Magnesian calcite cements andtheir diagenesis: dissolution and dolomitization,Mururoa Atoll: Sedimentology, v. 35, p. 821–841.

Al-Aasm, I.S., and J. Veizer, 1982, Chemical stabilizationof low-Mg calcites: an example of brachiopods: Jour-nal of Sedimentary Petrology, v. 52, p. 1101–1109.

Amthor, J.E., E.W. Mountjoy, and H.G. Machel, 1993,Subsurface dolomites in Upper Devonian Leduc For-mation buildups, central part of Rimbey-Meadow-brook reef trend, Alberta, Canada: Bulletin of the Canadian Society of Petroleum Geology, v. 41,p. 164–185.

Aulstead, K.L., R.J. Spencer, and H.R. Krouse, 1988,Fluid inclusion and isotopic evidence on dolomiti-zation, Devonian of western Canada: Geochimica etCosmochimica Acta, v. 52, p. 1027–1035.

Austin, J.A., 1983, OBC 5-A: overthrusting in a deep-water carbonate terrane, in A.W. Bally, ed., Seismicexpression of structural styles: AAPG Studies inGeology Series 15, v. 3, p. 167–172.

Badiozamani, K., 1973, The dorag dolomitizationmodel—application to the Middle Ordovician ofWisconsin: Journal of Sedimentary Petrology, v. 43,p. 965–984.

Baker, P.A., and S. Burns, 1985, Occurrence and forma-tion of dolomite in organic-rich continental marginsediments: AAPG Bulletin, v. 69, p. 1917–1930.

Barrett, M.L., 1987, The dolomitization and diagenesisof the Jurassic Smackover Formation, southernAlabama: Ph.D. thesis, The Johns Hopkins Univer-sity, Baltimore, Maryland, 362 p.

Bethke, C.M., 1985, A numerical model of compaction-driven groundwater flow and heat transfer and itsapplication to the paleohydrology of intracratonicsedimentary basins: Journal of GeophysicalResearch, v. 90B, p. 6817–6828.

Blatt, H., 1982, Sedimentary petrology: San Francisco,Freeman & Co., 564 p.

Bosellini, A., 1989, Dynamics of Tethyan carbonate plat-forms, in P.D. Crevello, J.L. Wilson, J.F. Sarg, and J.F.Read, eds., Controls on carbonate platform and basindevelopment: SEPM Special Publication, p. 3–13.

Bosellini, A., D. Masetti, and M. Sarti, 1981, A Jurassic“Tongue of the Ocean” infilled with oolitic sands:the Belluno Trough, Venetian Alps, Italy: MarineGeology, v. 44, p. 59–95.

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 159

Page 173: Reservoir Quality Prediction in Sand and Carbonates

160 Zempolich and Hardie

Broomhall, R.W., and J.R. Allen, 1985, Regionalcaprock-destroying dolomite on the Middle Jurassicto Early Cretaceous Arabian shelf: Society ofPetroleum Engineers, SPE 13697, p. 157–163.

Carballo, J.D., L.S. Land, and D.L. Miser, 1987,Holocene dolomitization of supratidal sedimentsby active tidal pumping, Sugarloaf Key, Florida:Journal of Sedimentary Petrology, v. 57, p. 153–165.

Carson, B., E. Suess, and J.C. Strasser, 1990, Fluid flowand mass flux determinations at vent sites on theCascadia margin accretionary prism: Journal ofGeophysical Research, v. 95, p. 8891–8898.

Casati, P., and M. Tomai, 1969, Il Giurassico ed ilCretacio del versante settentrionale del Vallone Bel-lunese e del Gruppo del M. Brandol: Riv. ItalianaPaleontologia e Stratigrafia, v. 75, p. 205–341.

Cati, A., D. Sartorio, and S. Venturini, 1987, Carbonateplatforms in the subsurface of the northern Adriaticarea: Mem. Soc. Geol. It., v. 40, p. 295–308.

Cervato, C., 1990, Hydrothermal dolomitization ofJurassic–Cretaceous limestones in the southern Alps(Italy): relation to tectonics and volcanism: Geology,v. 18, p. 458–461.

Coniglio, M., R. Sherlock, A.E. Williams-Jones, K. Mid-dleton, and S.K. Frape, 1994, Burial and hydrother-mal diagenesis of Ordovician carbonates from theMichigan Basin, Ontario, Canada, in B. Purser, M.Tucker, and D. Zenger, eds., Dolomites—a volume inhonour of Dolomieu: International Association ofSedimentologists Special Publication 21, p. 231–254.

Dawans, J.M., and P.K. Swart, 1988, Textural and geo-chemical alternations in Late Cenozoic Bahamiandolomites: Sedimentology, v. 35, p. 385–403.

Ditty, P.S., C.J. Harmon, O.H. Pilkey, M.M. Ball, andE.S. Richardson, 1977, Mixed terrigenous carbonatesedimentation in the Hispaniola Caicos turbiditebasin: Marine Geology, v. 24, p. 1–20.

Dix, G.R., 1993, Patterns of burial- and tectonicallycontrolled dolomitization in an Upper Devonianfringing-reef complex: Leduc Formation, PeaceRiver Arch area, Alberta, Canada: Journal of Sedi-mentary Petrology, v. 63, p. 628–640.

Doglioni, C., 1990, The Venetian Alps thrust belt, in K.R.McClay, ed., Thrust tectonics: London, Chapmanand Hall, p. 319–324.

Elder, J.W., 1965, Physical processes in geothermal areas,in W.H.K. Lee, ed., Terrestrial heat flow: AmericanGeophysical Union Monograph Series No. 8, p. 211–239.

Elder, J.W., 1967, Steady free convection in a porousmedium heated from below: Journal of FluidMechanics, v. 27, p. 29–48.

Elder, J.W., 1977, Thermal convection: Journal of theGeological Society of London, v. 133, p. 292–309.

Gaetani, M., and F. Jadoul, 1979, The structure of theBergamasc Alps: Rend. Acc. Naz. Lincei, v. 66, no. 5,p. 411–416.

Garven, G., 1985, The role of regional fluid flow in thegenesis of the Pine Point deposit, Western CanadaSedimentary Basin: Economic Geology, v. 80, p. 307–324.

Garven, G., and R.A. Freeze, 1984, Theoretical analysisof the role of groundwater flow in the genesis ofstratabound ore deposits 2: quantitatve results:American Journal of Science, v. 284, p. 1125–1174.

Ge, S., and G. Garven, 1989, Tectonically induced tran-sient groundwater flow in foreland basin: Int. A.G.U.Monograph Series, No. 48, I.U.G.G., v. 3, p. 145–157.

Gregg, J.M., 1985, Regional epigenetic dolomitizationin the Bonneterre dolomite (Cambrian), southernMissouri: Geology, v. 13, p. 503–506.

Gregg, J.M., and K.L. Shelton, 1989, Minor- and trace-element distributions in the Bonneterre Dolomite(Cambrian), southeast Missouri: evidence for possiblemulti-basin fluid sources and pathways during lead-zinc mineralization: Geological Society of AmericaBulletin, v. 101, p. 221–230.

Hanshaw, B.B., and W. Back, 1980, Chemical mass-wasting of the northern Yucatan Peninsula bygroundwater dissolution: Geology, v. 8, p. 222–224.

Hanshaw, B.B., W. Back, and R.G. Deike, 1971, A geo-chemical hypothesis for dolomitization by ground-water: Economic Geology, v. 66, p. 710–724.

Hardie, L.A., 1987, Dolomitization: a critical review ofsome current views: Journal of Sedimentary Petrol-ogy, v. 57, p. 166–183.

Jodry, R.L., 1969, Growth and dolomitization of Silurianreefs, St. Clair County, Michigan: AAPG Bulletin, v. 53, p. 957–981.

Kaufman, J., 1994, Numerical models of fluid flow incarbonate platforms: implications for dolomitization:Journal of Sedimentary Research, v. A64, p. 128–139.

Kohout, F.A., H.R. Henry, and J.E. Banks, 1977,Hydrogeology related to geothermal conditions ofthe Floridan Plateau, in The geothermal nature ofthe Floridan Plateau: Florida Bureau of GeologySpecial Publication 21, p. 1–41.

Kulm, L.D., et al., 1986, Oregon subduction zone: Vent-ing, fauna, and carbonates: Science, v. 231, p. 561–566.

Kupecz, J.A., and L.A. Land, 1991, Late-stage dolomiti-zation of the Lower Ordovician Ellenberger Group,West Texas: Journal of Sedimentary Petrology, v. 61,p. 551–574.

Kupecz, J.A., C. Kerans, and L.S. Land, 1988, Discus-sion: Deep-burial dolomitization in the OrdovicianEllenberger Group Carbonates, West Texas andSoutheastern New Mexico: Journal of SedimentaryPetrology, p. 908–910.

Kupecz, J.A., and L.A. Land, 1994, Progressive recrystal-lization and stabilization of early-stage dolomite:Lower Ordovician Ellenberger Group, west Texas, inB. Purser, M. Tucker, and D. Zenger, eds.,Dolomites—a volume in honour of Dolomieu: Inter-national Association of Sedimentologists Special Pub-lication 21, p. 255–279.

Kupecz, J.A., I.P. Montañez, and G. Gao, 1993, Recrys-tallization of dolomite with time, in R. Rezak and D. Lavoie, eds., Carbonate microfabrics, frontiers in sedimentology: New York, Springer-Verlag, p. 187–194.

Land, L.S., 1985, The origin of massive dolomite: Jour-nal of Geological Education, v. 33, p. 112–125.

Page 174: Reservoir Quality Prediction in Sand and Carbonates

Lee, Y.I., and G.M. Friedman, 1987, Deep-burialdolomitization in the Ordovician Ellenberger Groupcarbonates, West Texas and southeastern New Mex-ico: Journal of Sedimentary Petrology, v. 57, no. 3, p. 544–557.

Lee, Y.I., and G.M. Friedman, 1988, Reply: deep-burialdolomitization in the Ordovician Ellenberger Groupcarbonates, West Texas and southeastern New Mex-ico: Journal of Sedimentary Petrology, v. 58, p. 910–913.

Lucia, F.J., 1995, Rock fabric/petrophysical classificationof carbonate pore space for reservoir characterization:AAPG Bulletin, v. 79, p. 1275–1300.

Lumsden, D.N., 1985, Secular variations in dolomiteabundance in deep marine sediments: Geology, v. 13,p. 766–769.

Lumsden, D.N., 1988, Characteristics of deep-marinedolomite: Journal of Sedimentary Petrology, v. 58, p. 1023–1031.

Machel, H.G., and J.H. Anderson, 1989, Pervasive sub-surface dolomitization of the Nisku Formation ofCentral Alberta: Journal of Sedimentary Petrology,v. 59, p. 891–911.

Machel, H.G., and E.W. Mountjoy, 1986, Chemistryand environments of dolomitization—a reap-praisal: Earth Science Reviews, v. 23, p. 175–222.

Major, R.P., 1984, The Midway Atoll coral cap: meteoricdiagenesis, amplitude of sea level fluctuation, anddolomitization: Ph.D. thesis, Brown University,Providence, Rhode Island, 133 p.

Masetti, D., 1971, Sedimentologia e paleogeografia delGiurassico tra Brenta e Piave: Ph.D. thesis, Univer-sity of Ferrara, Italy, 104 p.

Masetti, D., and G. Bianchin, 1987, Geologia delGruppo della Schiara (Dolomiti Bellunesi). Suoinquadramento nella evoluzione giurassica delmargine orientale della Piattaforma di Trento:Mem. Ist. Geol. Min. Univ. Padova v. 39, p. 187–212.

Massari, F., P. Grandesso, C. Stefani, and P.G. Job-straibizer, 1986, A small polyhistory foreland basinevolving in a context of oblique convergence: theVenetian basin (Chattian to Recent, Southern Alps,Italy): International Association of SedimentologistsSpecial Publication 8, p. 141–168.

Mattes, B.W., and E.W. Mountjoy, 1980, Burial dolomi-tization of the Upper Devonian Miette buildup,Jasper National Park, Alberta: SEPM Special Publi-cation 28, p. 259–297.

McKenzie, J., 1981, Holocene dolomitization of cal-cium carbonate sediments from the coastal sabkhasof Abu Dhabi, U.A.E.: a stable isotope study: Jour-nal of Geology, v. 89, p. 185–198.

Miller, J.K., and R.L. Folk, 1994, Petrographic, geochemi-cal and structural constraints on the timing and distri-bution of postlithification dolomite in the RhaetianPortoro (“Calcare Nero”) of the Portovenere area, LaSpezia, Italy, in B. Purser, M. Tucker, and D. Zenger,eds., Dolomites—a volume in honour of Dolomieu:International Association of Sedimentologists SpecialPublication 21, p. 187–202.

Montañez, I.P., 1994, Late diagenetic dolomitization ofLower Ordovician, Upper Knox carbonates: arecord of the hydrodynamic evolution of the South-ern Appalachian Basin: AAPG Bulletin, v. 78, p. 1210–1239.

Montañez, I.P., and J.F. Read, 1992, Fluid-rock interac-tion history during stabilization of early dolomites,Upper Knox Group (Lower Ordovician), U.S.Appalachians: Journal of Sedimentary Petrology, v. 62, p. 753–778.

Morrow, D.W., 1982a, Diagenesis 1. Dolomite—Part 1:the chemistry of dolomitization and dolomite pre-cipitation: Geoscience Canada, v. 9, p. 5–13.

Morrow, D.W., 1982b, Diagenesis 2. Dolomite—Part 2:dolomitization models and ancient dolostones:Geoscience Canada, v. 9, p. 95–107.

Mountjoy, E.W., and J.E. Amthor, 1994, Has burialdolomitization come of age? Some answers from theWestern Canada Sedimentary Basin, in B. Purser, M.Tucker, and D. Zenger, eds., Dolomites—a volume inhonour of Dolomieu: International Association ofSedimentologists Special Publication 21, p. 203–229.

Mountjoy, E.W., and M.K. Halim-Dihardja, 1991, Multi-ple phase fracture and fault-controlled burial dolomi-tization, Upper Devonian Wabamun Group, Alberta:Journal of Sedimentary Petrology, v. 61, p. 590–612.

Mullins, H.T., N. Breen, J. Dolan, R.W. Wellner, J.L.Petruccione, M. Gaylord, B. Andersen, A.J. Melillo,A.D. Jurgens, and D. Orange, 1992, Carbonate plat-forms along the southeast Bahamas–Hispaniola col-lision zone: Marine Geology, v. 105, p. 169–209.

Mullins, H.T., S.W. Wise, L.S. Land, D.I. Siegel, P.M.Masters, E.G. Hinchey, and K.R. Price, 1985, Authi-genic dolomite in Bahamian periplatform slope sed-iment: Geology, v. 13, p. 292–295.

Prouty, C.E., 1983, The tectonic development of theMichigan Basin intrastructures, in R.E. Kimmel, ed.,Tectonics, structure, and karst in northern LowerMichigan: Michigan Basin Geological Society 1983Field Conference, p. 36–81.

Ritger, S., B. Carson, and E. Suess, 1987, Methane-derived authigenic carbonates formed by subduc-tion-induced pore water expulsion along theOregon/Washington margin: Geological Society ofAmerica Bulletin, v. 98, p. 147–156.

Ruppel, S.C., and H.S. Cander, 1988, Dolomitization ofshallow-water carbonates by seawater and sea-water-derived brines: San Andres Formation(Guadalupian), West Texas, in V. Shukla and P.A.Baker, eds., Sedimentology and geochemistry ofdolostones: SEPM Special Publication 43, p. 245–262.

Saller, A.H., 1984, Petrologic and geologic constraintson the origin of subsurface dolomite, EnewetakAtoll: an example of dolomitization by normal sea-water: Geology, v. 12, p. 217–220.

Sample, J.C., M.R. Reid, H.J. Tobin, and J.C. Moore, 1993,Carbonate cements indicate channeled fluid flowalong a zone of vertical faults at the deformation frontof the Cascadia accretionary wedge: Geology, v. 21, p. 507–510.

Geometry of Dolomite Bodies Within Deep-Water Resedimented Oolite of the Middle Jurassic Vajont Limestone 161

Page 175: Reservoir Quality Prediction in Sand and Carbonates

162 Zempolich and Hardie

Sandford, W.E., 1987, Assessing the potential for cal-cite dissolution in coastal saltwater mixing zones:Ph.D. thesis, Pennsylvania State University, StateCollege, Pennsylvania, 103 p.

Schlager, W., and N.P. James, 1978, Low-magnesian cal-cite limestones forming at the deep-sea floor, Tongueof the Ocean, Bahamas: Sedimentology, v. 25, p. 675–702.

Sears, S.O., and F.J. Lucia, 1980, Dolomitization ofnorthern Michigan Niagara reefs by brine refluxionand freshwater/seawater mixing, in D.H. Zenger,J.B. Dunham, and R.L. Ethington, eds., Conceptsand models of dolomitization: SEPM Special Publi-cation 28, p. 215–235.

Simms, M., 1984, Dolomitization by groundwater-flowsystems in carbonate platforms: Transactions of theGulf Coast Association of Geological Societies, v. 34,p. 411–420.

Sun, S.Q., 1995, Dolomite reservoirs: porosity evolutionand reservoir characteristics: AAPG Bulletin, v. 79,p. 186–204.

Tobin, H.J., J.C. Moore, M.E. MacKay, D.L. Orange, andL.D. Kulm, 1993, Fluid flow along a strike-slip fault atthe toe of the Oregon accretionary prism: implica-tions for the geometry of frontal accretion: GeologicalSociety of America Bulletin, v. 105, p. 569–582.

Trevisani, E., 1991, Il Toarciano-Aaleniano nei settoricentro-orientali della Piattaforma di Trento (PrealpiVenete): Riv. Italiana Paleontologia e Stratigrafia, v. 97, p. 99–124.

Vahrenkamp, V.C., and P.K. Swart, 1990, New distri-bution coefficient for the incorporation of strontiuminto dolomite and its implications for the formationof ancient dolomites: Geology, v. 18, p. 387–391.

van Tuyl, F.M., 1916, The origin of dolomite: IowaGeological Survey Annual Report, v. 25, p. 251–422.

Weber, J.N., 1964, Trace element composition of dolo-stones and dolomites and its bearing on thedolomite problem: Geochimica et CosmochimicaActa, v. 28, p. 1817–1868.

Weissert, H.J., and D. Bernoulli, 1985, A transformmargin in the Mesozoic Tethys: evidence from the Swiss Alps: Geologische Rundschau, v. 73, p. 665–679.

Weyl, P.K., 1960, Porosity through dolomitization:conservation-of-mass requirements: Journal of Sed-imentary Petrology, v. 30, p. 85–90.

Whitaker, F.F., P.L. Smart, V.C. Vahrenkamp, H.Nicholson, and R.A. Wogelius, 1994, Dolomitiza-tion by near-normal seawater? Field evidence fromthe Bahamas, in B. Purser, M. Tucker, and D.Zenger, eds., Dolomites—a volume in honour ofDolomieu: International Association of Sedimentol-ogists Special Publication 21, p. 111–132.

Wilkinson, B.H., and T.J. Algeo, 1989, Sedimentarycarbonate record of calcium-magnesium cycling:American Journal of Science, v. 289, p. 1158–1194.

Wilson, E.N., 1989, Dolomitization of the TriassicLatemar buildup, northern Italy: Ph.D. thesis, TheJohns Hopkins University, Baltimore, Maryland,272 p.

Wilson, E.N., L.A. Hardie, and O.M. Phillips, 1990,Dolomitization front geometry, fluid flow patterns,and the origin of massive dolomite: the TriassicLatemar buildup, Northern Italy: American Journalof Science, v. 290, p. 741–796.

Yao, Q., and R.V. Demicco, 1995, Paleoflow patterns ofdolomitizing fluids and paleohydrology of thesouthern Canadian Rocky Mountains: evidencefrom dolomite geometry and numerical modeling:Geology, v. 23, p. 791–794.

Zempolich, W.G., 1993, The drowning succession inJurassic carbonates of the Venetian Alps, Italy: arecord of supercontinent breakup, gradual eustaticrise, and eutrophication of shallow-water environ-ments, in R.G. Loucks and J.F. Sarg, eds., Carbonatesequence stratigraphy: recent developments andapplications: AAPG Memoir 57, p. 63–105.

Zempolich, W.G., 1995, Deposition, early diagenesis,and late dolomitization of deepwater resedimentedoolite: the Middle Jurassic Vajont limestone of theVenetian Alps, Italy: Ph.D. thesis, The Johns HopkinsUniversity, Baltimore, Maryland, 659 p.

Zempolich, W.G., and L.A. Hardie, 1991a, Massiveburial dolomitization: the Jurassic Vajont oolite ofnortheast Italy, in A. Bosellini, R. Brandner, E.Flügel, B. Purser, W. Schlager, M. Tucker, and D.Zenger, eds., Dolomieu Conference on CarbonatePlatforms and Dolomitization Abstracts: Ortisei,Italy, p. 298.

Zempolich, W.G., and L.A. Hardie, 1991b, Massive bur-ial dolomitization: the Jurassic Vajont oolite of north-east Italy (abs.): Geological Society of America,Abstracts with Programs, p. 411.

Zenger, D.H., 1976, Dolomitization and dolomite“dikes” in the Wyman Formation (Precambrian),northeastern Inyo Mountains, California: Journal ofPaleontology, v. 46, p. 457–462.

Zenger, D.H., 1983, Burial dolomitization in the LostBurro Formation (Devonian) East-Central Califor-nia, and the significance of late diagenetic dolomiti-zation: Geology, v. 11, p. 519–522.

Zenger, D.H., and J.B. Dunham, 1988, Dolomitization ofSiluro–Devonian limestones in a deep core (5350 m),southeastern New Mexico: SEPM Special Publication43, p. 161–173.

Page 176: Reservoir Quality Prediction in Sand and Carbonates

163

Chapter 11

Poroperm Prediction for Wildcat ExplorationProspects: Miocene Epoch, Southern Red Sea

Jon G. Gluyas1

BP Exploración de Venezuela S.A., Edificio Centro Seguros de Sud AmericaCaracas, Venezuela

Trevor WittonBP Exploration

London, England, United Kingdom

ABSTRACT

Prior to BP Exploration’s drilling the well Antufash-1 in the Yemeni watersof the Southern Red Sea, reservoir quality was estimated to be poor; it wasdry, plugged, and abandoned. The Miocene sandstones encountered weretight, with a mean porosity of 4% in the cored section and a permeability ofonly 0.07 md.

The prediction of low quality for the reservoir section of Antufash-1 wasbased on very few core analysis data. The diagenetic history of potentialreservoir sands in the Antufash acreage was calculated from data on depthto prospect, burial and thermal history of the area, reservoir sand prove-nance, and depositional environment.

An initial assessment, using limited local well data, led to the conclusionthat only at depths <0.5 km was it reasonable to expect high reservoir quality(>100 md). However, at depths >1.5 km, permeability was likely to be as lowas 10 md. Throughout this depth range, the chances of halite cementationwere also reasoned to be high. The rapid deterioration of reservoir qualitywith depth was attributed to the instability of the original volcaniclastic detri-tus. Such detritus was predicted to have converted to a mixture of zeolites andsmectitic clay soon after deposition. The reactivity of the assemblage was alsopredicted to have been exaggerated by the high thermal gradients in the area.

The recommendation was to avoid large parts of the license area known tohave received input of volcaniclastic sediment, and to develop prospects inthe few areas thought to have had arkosic sand input. These sands, it wasreasoned, would suffer less degradation of reservoir quality. The Antufash-1well successfully proved the existence of such arkosic sands in the basin, andtheir diagenetic history was as predicted. Unfortunately, the sandstoneswere tight. Halite cement filled, as predicted, all remaining porosity.

1Present address: Monument Oil and Gas plc, London, United Kingdom.

Gluyas, J.G., and T. Witton, 1997, Poroperm predic-tion for wildcat exploration prospects: MioceneEpoch, Southern Red Sea, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality pre-diction in sandstones and carbonates: AAPGMemoir 69, p. 163–176.

Page 177: Reservoir Quality Prediction in Sand and Carbonates

164 Gluyas and Witton

INTRODUCTION

In 1990, BP Exploration was awarded a 100% work-ing interest in the Antufash Concession, located off thewest coast of Yemen in the Southern Red Sea (Figure 1).In 1992 the commitment well Antufash-1 was drilled inthe northern part of the license. The well results demon-strated that the sandstones encountered were fullycemented, and only minor gas shows were recorded.

Earlier exploration drilling elsewhere in the South-ern Red Sea had resulted in a few very minor oil andcondensate discoveries. The general perception wasthat trap risk was relatively low, but significant riskswere associated with oil source and reservoir. Indeed,even if a working source were present, oil charge waspredicted to be limited due to the small prospectdrainage areas. However, the quality of the likelyreservoir interval remained the key risk.

The few sandstones that had been encounteredwere of very low permeability. The aim of our studywas to determine, ahead of the drill bit, the controls onreservoir quality, and to develop a methodology foreither mapping reservoir quality prior to drilling orestimating reservoir quality at the prospect locationprior to drilling.

DATABASE

Most of the reservoir quality data available at thetime of prospect evaluation were limited to qualitativedescriptions of cuttings, petrographic point-count datafor cuttings, and a few petrographic descriptions ofcore chips from wells scattered across the Red Sea. Afew scanning electron microscope (SEM) photomicro-graphs were available for some of the core chips.

Core analysis data (porosity and permeability) wereavailable for two old wells in the Antufash license (AlMeethag 1 and Al Meethag 2; Table 1).

GEOLOGICAL BACKGROUND

Basin Development

The Red Sea occupies a long (2000 km) linear rift oflate Oligocene age 180–50 km wide (Hughes et al.,1991). The conjugate margins are bounded by a seriesof large fault terraces with ≤2500 m of relief. Withinthe sea itself, there are three physiographic elements:(1) shallow shelfal areas, narrow north of 21°N butwider to the south; (2) a main trough 600–1000 m deepoccupying most of the sea area to the north of 21°N; and(3) a narrow axial trough, ~2000 m deep and 5–30 kmwide (Coleman, 1993).

The crust beneath the axial trough has been deter-mined as oceanic on the basis of magnetic anomalies(Girdler and Styles, 1978), with the oldest crust thoughtto have been formed ~5 Ma. There remains muchdoubt as to the nature of the crust beneath the shelfalareas. It may be oceanic crust formed during the Oligo–Miocene (Hall, 1989), thinned continental crust (Egloffet al., 1991), or a bit of both (Cochran, 1983). Sea-floorspreading is still active within the Red Sea area.

The thin continental crust and active sea-floorspreading has resulted in heat flows, typically >200mWm2 (megawatts per square meter) in the axialtrough (Coleman, 1993), and almost everywheregreater than the world average of 55 (mWm2). As aconsequence, thermal gradients are also high, from~73°C km–1 at the basin center to ~45°C km–1 at thebasin margin. Volcanism associated with the riftingcontinued throughout the Miocene (Davison andRex, 1980).

Stratigraphy

Four megasequences have been identified (Figure 2;Hughes and Beydoun, 1992; Mitchell et al., 1992). These,with their approximate ages, are described below.

Prerift Pre-Late Oligocene, 26 Ma or Older

The oldest megasequence comprises minormarine deposits generated by occasional flooding bythe Indian Ocean into the incipient rift in the Meso-zoic. The Oligocene is dominated by regional floodvolcanism.

BP Antufash License

Antufash-1

Al Meethag 1 (W1)

Al Meethag 2 (W2)

Wadi Mawr

YEMEN

YEMEN

RED SEA

42 Eo

15 No

16 No

43 Eo

Figure 1. Location map for the Antufash license inthe Yemeni Red Sea.

Page 178: Reservoir Quality Prediction in Sand and Carbonates

Synrift Early to Middle Miocene, 26–16 Ma (“Infra-Evaporite”)

In quick succession, the rift phase was typified byuplift, high rate of extension, and subsidence as Arabiaand Africa separated. Transgression occurred from thenorth, and widespread marine conditions were estab-lished. By the middle Miocene, the environmentremained shallow marine, but water circulation wasrestricted.

Postrift Middle to Late Miocene, 16–5 Ma (“Evaporite”)

Following development of a silled basin, thickevaporite deposits were developed during lowstanddrawdown. Uplift of the rift shoulders resulted indeposition of thick clastic wedges along the basin mar-gin. Intermittent connection with the Indian Oceanand periods of anoxia led to the development of poten-tial petroleum source rocks (El-Anbaawy et al., 1992;Cole et al., 1995). The massive salt at the base of thismegasequence also began to move at this time due tothe gravity loading in the coastal areas (Heaton et al.,1993; Davidson et al., 1994, 1995). Cyclic deposition of

source reservoir and seal parasequences occurred inthe salt withdrawal basins.

Axial Rift Pliocene–Recent, 5 Ma–Present (“Supra-Evaporite”)

Eustatic sea level fall accentuated erosion on thebasin margins. Spreading continued with rapid subsi-dence of the axial trough. New oceanic crust wasformed in the south, and open marine conditions wereestablished with the Indian Ocean; major carbonatedeposition occurred. The continued clastic depositionat the basin margins and the resultant salt movementaccentuated earlier developed structural traps.

Reservoir Development

A combination of seismic reflection data mappingand information obtained from existing wells revealedthat reservoir potential was likely to be best developedwithin upper Miocene highstand progradational sys-tems and associated lowstand systems tracts (Crossleyet al., 1992). In both tracts, basin-fringing alluvial/flu-vial systems were predicted to be the most likely reser-voirs. Some marine influence was likely to haveoccurred during maximum flood.

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 165

Table 1. Core Porosity and Permeability Data for Regional Wells in the Antufash License.

Average Mean Mean Depth Number Porosity 1σ Permeability 1σ

Well (m subsea) of Plugs (%) (%) (md) (md)

Al Meethag 1 1417 51 28.5 5.3 34.7 77.0(W1)

Al Meethag 2 1540 29 21.6 3.7 11.5 19.0(W2)

CHRONOSTRATIGRAPHY LITHOSTRATIGRAPHY GLOBAL SEQUENCESTRATIGRAPHY

MIO

CE

NE

PLIO-CENE

PLEISTOCENE

OLIGO-CENE

U

L

U

M

L

HOLOCENECALABRIAN -MILAZZIAN

PIACENZIAN

ZANCLEAN

MESSINIAN

TORTONIAN

SERRAVALLIAN

LANGHIAN

BURDIGALIAN

AQUITANIAN

SERIES STAGES SOUTHERNRED SEA

SUPRA-EVAPORITE

EVAPORITE

INFRA-EVAPORITE

MARINE SEDIMENTS &FLOOD VOLCANICS

RELATIVE CHANGEOF COASTAL ONLAP

SEQUENCEB'NDARY AGE

LANDWARD BASINWARD0.81.62.43.04.2

8.0

5.66.3

10.6

12.513.6

15.616.517.5

2122

24

Figure 2. Stratigraphy of theRed Sea area (R. Jones, 1994,personal communication).

Page 179: Reservoir Quality Prediction in Sand and Carbonates

166 Gluyas and Witton

Sand Provenance/Composition

The last major uplift in the Red Sea area began dur-ing the middle Miocene and continues today (David-son et al., 1994). As a consequence, the present-daygeological maps of the circum Red Sea area are takento represent the potential provenance area for UpperMiocene sediments (Sudan, 1963; U.S. Geological Sur-vey, 1963; Kazmin, 1973; Merla, 1979). In broad termsthere were two very different provenance terrains dur-ing the mid-Miocene:

• Pre-Cambrian acid and acid-intermediate meta-morphic and igneous granites and gneisses withminor Jurassic and Cretaceous sandstones (Taw-ila Formation).

• Pre-Cambrian basic metamorphic and igneousrocks, Oligo–Miocene volcanics, and minor Juras-sic limestones.

The following depositional sand compositions wereestimated using descriptions of SEM preparationsfrom core and cuttings obtained from the two AlMeethag wells (Figure 1).

These wells showed basic/volcanic-derived andacid-derived deposits as follows: quartz 30 ± 20% and50 ± 20%, respectively; feldspar 30 ± 20% and 30 ± 20%,respectively; various rock fragments 40 ± 20% and20%, respectively.

The rock fragments include volcaniclastic grains,mafic mineral grains, and a little glauconite.

In the Antufash acreage, as evidenced by the AlMeethag wells, much of the provenance appears tohave been from the basic metamorphics and volcanics.Only in the area west of Wadi Mawr (Figure 1) is thisbasic/volcanic input likely to have been diluted. Thiswadi drains along a Jurassic transfer fault and taps intoan area that may have shed large quantities of arkosicTawila Sandstone during the middle Miocene.

EVALUATION OF RESERVOIR QUALITY

Method

Two approaches were used to calculate reservoirquality for the Miocene sandstones of the SouthernRed Sea. First, the limited poroperm data that did existwere analyzed in terms of the controls on porosity andpermeability, using the methodology of Cade et al.(1994). The results from this analysis were comparedwith the qualitative petrographic descriptions.

Second, reservoir quality data were evaluated usingthe broad geological data available for the area. Themethods for porosity and permeability synthesis aregiven below.

Correct prediction of porosity requires that the vol-ume losses due to compaction and cementation arequantified. Permeability prediction further requiresknowledge of the grain size and sorting characteris-tics, cement types, and their distribution.

Porosity loss due to compaction was calculated usingthe methodology of Gluyas and Cade (this volume).Cement types and volumes were calculated on the basis

of a BP Exploration in-house regional diagenesis study(Primmer, 1993; Primmer et al., this volume), in whichthe links between sand mineralogy at deposition, depo-sitional environment, burial, and thermal history werequantified.

Grain size and sorting data were adopted from theexisting well information. Permeabilities were calcu-lated from the cement data and estimates of grain sizeand sorting, using the sphere pack modeling approachof Cade et al. (1994).

Data Analysis

The reservoir intervals of the two Al Meethag wellscontain quartz-poor, feldspar- and volcaniclastic-rich,fine- to medium-grained sandstones. Their diagenetichistory is complex, with calcite, dolomite, chlorite,smectite, zeolite, quartz, illite, and halite cements.Given that the sediments are at most 15 m.y. old andeven now buried only to 1.5–1.7 km, all of theseprocesses must have occurred in a short geologicaltime and at shallow depth.

An attempt to construct an empirical porosity depthplot proved futile. The problems encountered are illus-trated in Figures 3 and 4. In short, there are too fewdata from which to draw any valid conclusions as tohow, or if, porosity varies with depth in this basin.

Porosity and permeability data from conventionalcore analysis for the two Al Meethag wells are plottedin Figure 5. Plotted on the same graph are modeledcurves for the porosity-to-permeability relationship insimilar grain size (fine- to medium-grained) clean,compacted, and/or quartz cemented sandstones(Evans et al., this volume). Most of the data from thetwo wells describe two distinctly different prolateclusters of reasonably similar permeability range butsignificantly different porosity range. The outliers tothese two trends are medium/coarse grained sand-stones, carbonate cemented sandstones, and, in oneinstance, a fractured core analysis plug.

The poroperm data for both wells lie well below themodeled clean sand lines. The steep porosity-to-permeability gradient is indicative of a sand with alarge proportion of poorly interconnected porosity:either intragranular secondary porosity or micropo-rosity trapped between clay fibers and plates. The sim-ilarity of poroperm gradients in the two wells was takento indicate that the process controlling permeabilityevolution in both was similar. This relationship didnot hold for porosity. The inferred importance of clayin controlling the permeability of these sandstones isfully supported by the petrographic descriptions.

In order to explain the porosity difference betweenthe wells, a process is needed to reduce porosity withonly a minor (relative to the clays) effect on permeabil-ity. At the high porosities seen in these cores, twoprocesses could have been responsible: compactionand/or syntaxial quartz precipitation (Cade et al., 1994).

There is insufficient difference in burial depth of thetwo sandstones to account for the porosity difference interms of compaction alone, even when the 13 MPaoverpressure in well W1 is taken into account (Robin-son and Gluyas, 1992). It is possible that quartz

Page 180: Reservoir Quality Prediction in Sand and Carbonates

cementation may account for much of the difference.This suggestion is supported by the qualitativedescriptions of the petrography of the sandstonesfrom the two wells. Quartz cement was describedfrom W2 but not from W1. Modeling data (see thefollowing section) also lend some support to thissuggestion.

No equivalent quantitative reservoir quality datawere available for the sandstones derived from thePre-Cambrian acid igneous and gneiss terrains.

Data Synthesis—Modeled Poroperm Evolution

The following criteria were used to construct asemiquantitative diagenetic history for the Miocenesandstones (Figures 6, 7).

Volcaniclastic Sandstones

• Volcaniclastic sandstones are likely to react in situat temperatures below 25°C to produce aluminiumand iron smectites, zeolites (clinoptilolite), andchlorite. By 75°C, the same assemblage can furtherreact to produce higher temperature zeolites at theexpense of aluminum smectite. At 100°C, lau-monite is likely to be the stable zeolite alongsidealbite and quartz and the persistent chlorite (Blochand Helmold, 1995; Primmer et al., this volume).

• At temperatures >70°C, burial rates exceeding~100 m.y.–1 (meters per million years), and heat-ing rates exceeding 2°C m.y.–1, quartz is likely tobe an important cement phase during open sys-tem diagenesis (Gluyas et al., 1993).

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 167

Al Meethag 1

Al Meethag 1 overpressure

corrected

Al Meethag 2

-2000

-1800

-1600

-1400

-1200

-1000

-800

-600

-400

-200

0

0 5 10 15 20 25 30 35 40

Porosity (%)

Dep

th (

m, s

ubse

a)

Figure 3. Porosity depth plot for cored intervals from Antufash License. Al Meethag 1 is plotted twice, at itscurrent burial depth and at its hydrostatic equivalent burial depth. The Miocene sands in Al Meethag 1 areoverpressured by 9 MPa (1300 psi); 1 MPa is ~80 m of burial in a hydrostatic system at these burial depths(Gluyas and Cade, this volume). Data are averages for wells; individual plug data are plotted in Figure 5.Circle = Al Meethag 1; square = Al Meethag 2; diamond = Al Meethag 2 overpressure corrected.

Figure 4. Porosity–depth plot for log data from intervals in Al Meethag 1. Porosity and shale percentages werecalculated from a combination of neutron density and resistivity logs.

porosity (%)

buria

l dep

th (

m)

-1800

-1700

-1600

-1500

-1400

-1300

-1200

-1100

-1000

-900

-800

0 10 20 30 40

0-10% shale

11-20% shale

21-50% shale

50-75% shale

>75% shale

core

Page 181: Reservoir Quality Prediction in Sand and Carbonates

168 Gluyas and Witton

• There are sufficient components for illite to form,although significant quantities are unlikely toexist at temperatures below about 100°C (Small etal., 1992).

• In a depositional system containing some marineinfluence, a little early diagenetic carbonate is to beexpected (Bjørkum and Walderhaug, 1990); someof this cement is likely to have been dissolved andreprecipitated during the later stages of diagene-sis. Some decarboxylation carbonate may havealso been added (Gluyas and Coleman, 1992).

• In sequences interbedded with evaporites, there isa possibility that any residual porosity will havebeen filled by halite and other evaporite minerals.This point is speculative. We do not yet haveinformation that would allow us to describe theprocess or timing of such cementation.

• Finally, we made the assumption that of the compo-nents required for silicate mineral cementation, onlysilica is likely to have been imported to the sands inquantities large enough to appreciably affect poros-ity (Gluyas and Coleman, 1992). This point could beconsidered controversial given the current debate inthe literature with respect to the sources of silica forquartz cementation. However, we imply no scale oftransport here; import could mean derivation fromlocal silica sources, such as nearby pressure dissolu-tion seams, or more distant sourcing from unspeci-fied sources. Other elements such as potassium and

0.01

0.1

1

10

100

1000

10000

100000

0 5 10 15 20 25 30 35 40

Porosity (%)

Per

mea

bilit

y (m

d)

Figure 5. Porosity and permeability data from thecored intervals of Antufash License wells. AlMeethag 1 (blue) high porosity; Al Meethag 2(brown) lower porosity. The 900-md outlier is from afractured plug; the remaining outliers are from thin,medium-grained sandstones.

Figure 6. Synthesized diagenetic history for volcani-clastic sandstones in the Antufash License.

Figure 7. Synthesized diagenetic history for arkosicsandstones in the Antufash License.

15 10 5 0

Deposition

Carbonateprecipitation

Chloriteprecipitation

Compaction

Smectite &zeolite ppt.

Quartzprecipitation

Illite precipitation

Oil migration

Porosity &permeability

evolution

high

low

porosity

permeability

Ma

Deposition

Carbonateprecipitation

Compaction

Kaoliniteprecipitation

Quartzprecipitation

Illite precipitation

Oil migration

Porosity &permeability

evolution

high

low

porosity

permeability

Ma

and / or

15 10 5 0

Page 182: Reservoir Quality Prediction in Sand and Carbonates

aluminum are likely to have been supplied inter-nally (Gluyas and Leonard, 1995).

Arkosic SandstonesThe diagenesis of arkosic sandstones is likely to

have been very different (Figure 7). The most commonlow-temperature product is likely to have been kaolin-ite, precipitation of which could have accompaniedingress of undersaturated water of near-surface, mete-oric, or connate origin (Gluyas, 1985; Bjørkum et al.,1993). In an open system, quartz is likely to have pre-cipitated once the sandstones exceeded 70°C. By100°C, illite will have been the most likely clay phaseto precipitate (Small et al., 1992).

The presence of carbonate and evaporite cement islikely to be common to both the volcanic and arkosicsourced sandstones, since both cements would havebeen supplied from largely outwith the sandstone.

Estimating Porosity and Permeability—Antufash-1

The location for Antufash-1 is shown in Figure 1. Thearea was a poorly explored anticline/diapir fairwaycomprising upper-middle Miocene reservoirs. Theprospect lay above a well-defined NNW-SSE–trendingsalt-cored anticline with four-way dip closure through-out the Pliocene and Miocene sections. Multiple reser-voirs were expected to be present in transgressive sands

of the upper-middle Miocene sections. Four such reser-voirs were included in the volumetric calculations. Theseal was expected to be salt. Depth to crest of the upper-most prospective horizon was estimated at 850 m, andgross reservoir thickness was calculated at 450 m. The surface temperature is 25°C, and the present thermalgradient is 60°C km–1. The reservoir was expected to benormally pressured. The sands were estimated to be finegrained and well sorted. The sediment source area wasthought to be along Wadi Mawr (Figure 1), which drainsdominantly granitic terrain and is likely to have yieldedarkosic sands.

Modeled Porosity

Using the above criteria, the effects of compactionand quartz cementation were modeled. The likelyeffect of carbonate cement on bulk porosity wasassumed to be small, by analogy with the Al Meethagwells, while the potential for evaporite plugging ofporosity was estimated to be large.

The modeled porosity–depth curve for eitherarkosic or volcaniclastic sands in the Antufash acreageis shown in Figure 8. In order to generate such aporosity–depth curve, several simplifying assump-tions were made. Those that might have introduced asystematic error in the porosity estimate are:

• Formation of overpressuring during burial, lead-ing to a low porosity estimate

• Conversion of labile volcaniclastic grains to duc-tile “clay clasts,” which are more susceptible tocompaction than rigid grains, leading to an over-estimation of porosity

The sensitivity of the porosity estimate at 1 km bur-ial, errors in pressure, or ductile grain content areexamined in Table 2.

The porosity gradient associated with cementation,16 ± 5% km–1, is based on the empirical observationthat subregional porosity gradients resulting fromquartz cementation covary with thermal gradients(Rønnevik et al., 1983) (Table 3; Figure 9).

Modeled Permeability

In addition to the data required for porosity calcula-tion, data on grain size, sorting, and cement mineral-ogy were required for the permeability estimate. Grainsize data were taken from the Al Meethag wells and,for want of hard data, sorting was assumed to be mod-erate. Two cases were run for the cement mineralogy.For the volcaniclastic sands, a case based on pervasive,pore-lining smectite and/or zeolites was calculated,while the arkosic sand calculation was based on a porestructure comprising clean, “grain-lined” pores withrandomly scattered kaolinite-filled pores.

Modeled curves of the porosity to permeabilityrelationship for both the arkosic and volcaniclasticcases are shown in Figures 10 and 11.

Using the porosity/permeability relationshipsand porosity-to-depth relationships, it is possible todetermine depth equivalencies for the permeabilitycutoffs (100 md, 10 md) (Table 4).

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 169

Figure 8. Synthesized and simplified porosity–depthrelationship for all types of Miocene sandstones ofthe Antufash License. Curve C-C is a pure com-paction curve for a rigid grain hydrostatically pres-sured sandstone; curve Q-Q is the expected porositydecline of 16% km–1 hung from a depth equivalenceof ~70°C (800 m) at the time of silicate cementation(quartz, clays, and/or zeolites). Q+ and Q– are thepotential ±5% km–1 variance on the expected value.

Porosity (%)D

epth

(m

)

-2500

-2000

-1500

-1000

-500

0

0 10 20 30 40 50

C

C

Q

Q Q+Q-

Page 183: Reservoir Quality Prediction in Sand and Carbonates

170 Gluyas and Witton

Prospect-Specific Estimates of Porosity andPermeability

The estimated depth to top reservoir was 850 m. Atthis shallow depth, the risk on reservoir effectivenesswas low for arkosic and volcaniclastic sandstones.Both types of sandstones were likely to have porosities~36% and permeability >2 darcys.

ANTUFASH-1—WELL RESULTS

Antufash-1 was drilled to 2062 m in December 1992.The terminal depth was in middle Miocene halite andanhydrite, having penetrated Pleistocene, Pliocene,and upper Miocene sequences. Hydrocarbon gaspeaks were encountered while drilling through mud-stone intervals with source potential. However, otherthan these minor shows, the well was dry.

Thin sands were encountered in the upper part of themiddle Miocene section (1200 m), and thicker sand-stones were found at 1700 m within the middle Miocenesection. Both sequences were ~200 m deeper thanexpected. The sandstones at both 1200 m and 1700 mwere largely tight. Rig-site core examination led to theconclusion that much of the cement was halite. This waslater confirmed from detailed petrography on samplescut from the core without using water as a lubricant.The cored intervals were at hydrostatic pressure.

Cores 1 and 2, cut from the shallower sandstones at1200 m (Table 5), comprise a mixture of mudstones witha few fine- to medium-grained sandstone and disrup-tive, authigenic evaporite mineral layers and dikes.These intervals were interpreted to have been depositedin a marginal marine environment on the basis of sedi-mentary structures and the presence of fish debris.

Core 3, cut from deeper sandstones at 1700 m, com-prises sandstone (90%) and mudstones rich in organicmatter (10%). The sandstones occur as 1–1.75 mupward-fining, coarse to fine, cross-laminated sand-stones. Mud-clast lags are common in the basal partsof the beds. Planar bedded sandstones with currentripple tops are also present. The sediments of core 3are interpreted to have been deposited in high-energyfluvial channels and overbank areas.

Both the marine and fluvial sequences are compara-ble to those encountered in the Al Meethag wells.

Due to the extensive halite cementation, plugs forcore analysis and thin-section preparation were cut

Table 3. Porosity Gradients and Thermal Gradients—Sandstones Around the World.

Porosity Thermal Gradient Gradient

Sandstone Location (% km–1) (°C km–1) Reference

Brae North Sea 9 33 Gluyas, 1985(sandstone A)

Brent North Sea 7–8 30 Gluyas, 1985(sandstone B)

Stø Barents Sea 16 60 Rønnevik et al., 1983Sihapas Sumatra 20 60+ Gluyas and Oxtoby, 1995Garn Haltenbanken 8 30 Ehrenberg, 1990— E. Pacific 12.8 35 Bjørlykke et al., 1989— E. Pacific 8.5 25 Bjørlykke et al., 1989— Gulf Thailand 11 49 Bjørlykke et al., 1989San Joaquin California 6.4 35 Bjørlykke et al., 1989Frio Texas Gulf Coast 6.7 32.2 Loucks et al., 1984Frio Texas Gulf Coast 4.8 38.3 Loucks et al., 1984Jackson Texas Gulf Coast 7.5 20 Loucks et al., 1984Queen City Texas Gulf Coast 6.1 20 Loucks et al., 1984Mungaroo NW Shelf, Australia 5.0 18 Gluyas et al., 1993

Table 2. Sensitivity of Porosity to Compaction as aFunction of Burial Depth, Overpressure, andDuctile Grain Content.

Burial Ductile Depth Grain Overpressure Porosity(m subsea) Content (%) (MPa) (%)

1000 0 0 351000 10 0 321000 20 0 301000 30 0 291000 50 0 261000 0 1 361000 10 1 321000 20 1 311000 30 1 291000 50 1 271000 0 2 361000 10 2 331000 20 2 311000 30 2 301000 50 2 271000 0 5 381000 10 5 351000 20 5 331000 30 5 321000 50 5 301000 50 10 35

Page 184: Reservoir Quality Prediction in Sand and Carbonates

using diesel as a lubricant. A few of the plugs werecleaned of halite; their porosity and permeability weremeasured during and after the cleaning process. Fiveporosity and two permeability measurements weremade on cores 1 and 2; 36 pairs of measurements weremade on core 3. Unless otherwise stated, the remain-der of the discussion centers on core 3.

Before cleaning, the average porosity of the sand-stones was 1.5% and permeability was <1 md. Dataare presented in Table 6 for the five samples subjectto cleaning.

The sandstones of Antufash-1 are highly felds-pathic. The major portions are: quartz plus polycrys-talline quartz (42%), total feldspar (35%), and lithicfragments (22%). The feldspar is largely orthoclaseand extensively altered. Microcline is also abundant.The rock fragments are largely of igneous origin, com-posed of quartz, feldspar, and mica aggregates. Mostsamples contain a few percentages of metamorphicrock fragments and volcaniclastic rock fragments.

The sandstones contain an average of 38% diageneticcomponents. More than half of this cement is halite, andmuch of the remainder is dolomite. Some feldsparappears to have undergone a late diagenetic alteration tomica. A few percentages each of siderite, calcite, quartz,anhydrite, kaolinite, and feldspar are present with traceamounts of pyrite and anatase (Figure 12; Table 7).

Detrital Mineralogy

The sandstone mineralogy of Antufash-1 is withinthe range of mineralogy expected for a granitic source.This supports the hypothesis of a Wadi Mawr sourcefor the sediment.

The predicted composition/granitic source and Antu-fash-1 compositions of quartz are 50 ± 20% and 42%,respectively; feldspar, 30 ± 20% and 35%, respectively;and lithic fragments, 20 ± 20% and 23%, respectively.

Diagenetic Mineralogy

The predicted and actual mineral parageneses aresimilar (Table 8). Early diagenetic carbonate was fol-lowed by feldspar dissolution with quartz and kaoliniteprecipitation, with late diagenetic “mica” and more car-bonate. Halite was the last mineral to precipitate.

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 171

thermal gradient (°C km–1)

poro

sity

gra

dien

t (%

/km

)

0

5

10

15

20

25

0 20 40 60 80

Figure 9. Empirical relationship between porositygradient (due to quartz cementation) and thermalgradient for sandstones worldwide (Table 3).

Figure 10. Modeled porosity-to-permeability relationship for volcaniclastic sandstones of theAntufash License. C = compaction/quartz cementationcurve for fine (200 µm), moderately sorted sandstone. L = poroperm relationship for pore-lining clay cemented sandstone. Porosity of 36% isderived from Figure 8 (850 m burial). k = predictedpermeability. The system contains about 10% pore-lining clay.

Porosity (%)

0.1

1

10

100

1000

10000

100000

0 10 20 30 40

k

C

LPer

mea

bili

ty (

md

)

Porosity (%)

0.01

0.1

1

10

100

1000

10000

100000

0 10 20 30 40

C

F

k

Per

mea

bili

ty (

md

)

Figure 11. Modeled porosity-to-permeability relation-ship for arkosic sandstones of the Antufash License.C = compaction/quartz cementation curve for fine(200 µm), moderately sorted sandstone. F = poropermrelationship for pore-filling clay cemented sandstone.Porosity of 36% derived from Figure 8 (850 m burial).k = predicted permeability. The system containsabout 10% pore-filling clay.

Page 185: Reservoir Quality Prediction in Sand and Carbonates

172 Gluyas and Witton

Reservoir Quality

A correct prediction of diagenetic sequence is still along way from a quantitative prediction of porosityand permeability. Table 9 shows that the abundance ofquartz was overestimated, while that of early diage-netic carbonate (dolomite) was underestimated. Thelikely presence of pervasive halite was indicated, andthe reservoir effectiveness risked accordingly. How-ever, no attempt was made to quantify the halite vol-ume prior to drilling the well.

The predicted and actual poroperms are shown inFigure 13 and Table 9. In the very broadest of terms, theprediction of poroperm was perfect insofar as the prob-ability of halite plugging pores was estimated to behigh. On a halite-minus porosity basis, the predictionsfailed. For a prospect depth of ~1800 m, the estimatedporosity for a hydrostatically pressured, arkosic sandreservoir was ~20 ± 2.5%. This compares with 27 ± 3%in Antufash-1. Permeability for this same sandstone

was calculated at 100 md; this is less than the 2700 ±1900 md encountered in Antufash-1. If the comparisonof halite-minus porosity (predicted) vs. halite-minusporosity (from six cleaned plugs) is a valid (and it maynot be), our poroperm predictions were low.

There are a several reasons why predicted andactual halite-minus porosities differ: laboratory disso-lution of halite may not replicate the porosity prior tohalite cementation; the six samples selected for haliteremoval may not be representative of the sandstone;

Figure 12. Paragenetic sequence for the Miocenesandstones of Antufash-1.

Table 5. Cores Cut in Antufash-1.

m (BRT)* Top Base

Core 1 1196 1206.73Core 2 1211 1222.70Core 3 1761 1770.70

*BRT = below rotary table.

Table 6. Reservoir Quality of Sands After Partial and Complete Dissolution of Halite.

Porosity Permeability Porosity Permeability Porosity PermeabilityPlug (%) (md) (%) (md) (%) (md)Number Unclean Unclean Partly Clean Partly Clean Clean Clean

6 1.1 0.11 11.5 564 24.4 70913 0.6 0.17 10.4 646 25.5 153015 1.7 0.67 14.7 2040 28.6 508024 0.8 0.04 14.1 1930 29.7 326035 0.3 0.02 6.5 7.8 20.8 14Average 0.90 0.20 11.44 1038 25.8 2119st dev (1σ) 0.53 0.27 3.29 900 3.5 2052av (without plug 35) 1.05 0.25 12.68 1295 27.1 2645st dev (without plug 35) 0.48 0.29 2.06 799 2.5 1941

Deposition Today

Dolomiteprecipitation

Compaction

Kaoliniteprecipitation

Quartzprecipitation

Anhydriteprecipitation

Porosity &permeability

evolution

high

low

porosity

permeability

Time

Deposition, pyrite/ siderite ppt

Illitic clayprecipitation

Haliteprecipitation

Table 4. Depth and Porosity Criteria for Effective Reservoir (for Oil).

Porosity (%) Depth (mss)*Permeability Equivalent to Equivalent to

Sand Cutoff (md) Permeability Permeability

Arkosic 10 14 2200Volcaniclastic 10 30 1100Arkosic 100 20 1800Volcaniclastic 100 32 1000

* mss = meters subsea.

Page 186: Reservoir Quality Prediction in Sand and Carbonates

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 173

Tab

le 7

. Poi

nt-

Cou

nt M

iner

alog

y of

San

dst

ones

Fro

m C

ore

3 of

An

tufa

sh-1

.

Dep

th (m

KB

)*17

61.3

1762

.217

64.3

1764

.917

65.8

1766

.417

67.3

1767

.317

67.6

1768

.817

69.1

1769

.717

70.6

Gra

in s

ize

csL

msL

msU

msU

msU

msU

csU

csU

msU

msU

msU

msL

msL

Sor

tin

gm

ws

vws

ws

ws

mw

sw

sw

sm

ws

ws

ws

ws

ws

ws

Mon

oqu

artz

15.0

22.5

19.0

13.5

16.0

11.5

7.0

8.5

15.5

19.5

4.5

19.5

13.5

Pol

yqu

artz

12.5

9.5

7.0

15.5

10.0

10.5

10.5

17.0

12.5

10.5

6.0

5.5

8.0

Feld

spar

19.0

16.0

22.5

14.5

19.5

16.5

9.5

17.0

15.5

17.0

3.0

21.0

14.5

Ign

eou

s/m

eta

9.0

8.5

8.5

10.0

14.5

14.5

5.0

13.5

13.0

14.0

3.0

11.5

14.5

frag

men

tsV

olca

nic

1.

51.

00.

50.

51.

00.

50.

5—

1.0

1.0

—0.

51.

0fr

agm

ents

Sed

imen

tary

roc

kt

1.5

1.5

1.0

3.5

1.0

34.5

6.0

1.5

3.5

50.0

0.5

1.5

frag

men

tsH

eavy

tt

—0.

5t

tt

t—

t0.

5m

iner

als

Mic

a0.

53.

51.

51.

53.

02.

01.

51.

51.

5t

—3.

56.

0

Mat

rix

clay

0.5

0.5

t—

0.5

——

——

—0.

5—

—O

rgan

ic

—0

——

——

——

——

0.5

——

mat

ter

Qu

artz

3.

52.

51.

51.

00.

51.

0t

1.0

3.0

2.0

—5.

03.

0ce

men

tF'

elds

par

1.0

1.5

1.0

0.5

t1.

0t

1.0

1.0

0.5

—1.

00.

5ce

men

tH

alit

e27

.520

.522

.023

.021

.529

.06.

521

.522

.521

.51.

513

.518

.0

An

hyd

rite

0.5

1.5

4.5

tt

1.0

2.0

0.5

t0.

55.

51.

03.

0D

olom

ite

5.0

8.0

6.0

14.0

7.5

4.5

20.5

9.0

6.5

5.0

23.0

9.0

9.5

Kao

lin

ite

2.5

2.0

2.0

3.0

1.0

3.5

1.5

2.0

1.0

1.5

1.0

2.0

1.5

Illi

tic

clay

1.5

0.5

1.5

1.0

1.0

3.0

0.5

1.5

5.5

3.5

—6.

04.

5P

yrit

et

0.5

0.5

——

—0.

5t

—t

0.5

——

An

atas

e0.

5t

0.5

1.0

t0.

5—

t—

——

0.5

0.5

*mK

B =

met

ers

Kel

ly B

ushi

ng (a

mea

sure

of d

epth

from

the

rig

floo

r).

Page 187: Reservoir Quality Prediction in Sand and Carbonates

174 Gluyas and Witton

halite cementation in nature may not have occurredwhen predicted with respect to porosity evolution.

DISCUSSION

Estimates of reservoir porosity and permeability area fundamental part of a prospect evaluation. The mostcommonly used method of predicting a parametersuch as porosity (or permeability) is to take an existingdata analog, or perhaps a global data set, and plotporosity against something that correlates with it andis predictable. The first thing to attempt is a correlationwith depth. If a porosity-to-depth correlation is accept-able, the process may go no farther. However, if noother parameters are added in, such as “shalyness” orreservoir age or thermal maturity or ... (Schmoker andGautier, 1988).

The empirical approach can work well, but does,however, suffer from some drawbacks. The very natureof the approach allows prediction of the average.Anomalies, sandstones that are uncommonly porous atdepth or well-cemented at shallow depths, go unrecog-nized. More fundamentally, in the case of the Antufasharea, there were too few data on which to construct anysort of empirical porosity–depth plot. By consideringdiagenetic process, we were able to recognize that thetwo Al Meethag wells with their porous but imperme-able volcaniclastic sandstones could hardly be called

candidate reservoirs. The need to avoid volcaniclasticsandstones led to work directed at identifying adjacentareas onshore, which might have yielded a sand withgreater chemical stability. The perception of reservoireffectiveness as a function of sediment composition wasincorporated within the prospect-specific risks for theAntufash License prior to drilling.

The diagenetic sequence encountered in Antufash-1was close to that predicted, but mineral volume esti-mates were incorrect. Because almost all of the porosityin the Antufash sands was plugged by halite, whichcould have limited precipitation of other phases (noporosity left to fill), a quantitative comparison of pre-dicted vs. actual poroperm for Antufash-1 is difficult.

CONCLUSIONS

The approach used here for prediction of reservoirquality ahead of drilling Antufash-1 was an attempt atquantification of the sandstone reservoir quality as afunction of depositional characteristics and burial his-tory, conditioned to existing well data. Many assump-tions were made in order to generate the prediction ofreservoir quality. We have justified our assumptionson the basis of empirical observations.

The absolute numerical success of this approach forAntufash-1 is difficult to assess. However, the disciplineimposed on the poroperm prediction methodology by

Table 8. Predicted and Actual Diagenetic History for Antufash-1.

Diagenetic Predicted Composition Antufash-1 Sequence —Granitic source Composition

Earliest calcite precipitation dolomite precipitation

kaolinite ± quartz precipitation quartz, kaolinite,quartz, feldspar dissolution feldspar precipitation,

and dissolution

illite ± ferroan dolomite dolomite, and then anhydriteprecipitation + mica precipitation

Latest halite precipitation halite precipitation

Table 9. Quantitative Effect of Cementation on Porosity.

Diagenetic Predicted Abundance Antufash-1 Minerals —Granitic Source Abundance

Early diagenetic 10% reduction on net:gross, average 9%, carbonate no effect elsewhere range 2–23%

Quartz minor 2%, range 0.5–3.5%Kaolinite minor 2%, range 0–6%Illite/mica minor 2%, range 0–6%Late diagenetic minor minor

carbonateHalite not quantified 2%, range 1.5–25%

Page 188: Reservoir Quality Prediction in Sand and Carbonates

the need to understand how diagenetic processes werecontriving to reduce reservoir quality had an importantoutcome that would have been unavailable to a“guessed prediction” of reservoir quality. The well wasdrilled where the volcaniclastic component of the sandwas expected to be low. Thus, we created an interven-tion to mitigate against the high risk of encounteringporous but impermeable reservoir.

ACKNOWLEDGMENTS

The authors thank BP Exploration for allowing pub-lication of this paper. The paper was improved by thecomments and suggestions of Keith Myers and BobJones. We also thank David Lawrence and John Aggett,who logged the cores from Antufash-1, and subse-quently performed the petrographic analysis. To getthis far with poroperm prediction would not have beenpossible without the efforts of an outstanding “Reser-voir Quality Prediction” team at BP Research between1991 and 1994; we thank: Chris Cade, Shona Grant,Andrew Hogg, Mark Hopkins, Norman Oxtoby, TimPrimmer, Craig Smalley, Ed Warren, and Richard Wor-den. David G. Roberts and Richard Heaton are thankedfor their thorough reviews of this paper.

REFERENCES CITED

Bjørkum, P.A., R. Knarud, and M. Bergan, 1993, Howimportant is the late Cimmerian unconformity in

controlling the formation of kaolinite in sandstonesof the North Sea (examples from the Snorre field)?, inA.D. Harbury and A.G. Robinson, eds., Diagenesisand basin development: AAPG Studies in Geology36, p. 261–269.

Bjørkum, P.A., and O. Walderhaug, 1990, Geometricalarrangement of calcite cemented layers in shallowmarine sandstones: Earth-Science Reviews, v. 29, p. 145–161.

Bjørlykke, K., M. Ramm, and G. Saigal, 1989, Sandstonediagenesis and porosity modification during basinevolution: Geologische Rundschau 78, p. 243–268.

Bloch, S., and K.P. Helmold, 1995, Approaches topredicting reservoir quality in sandstones: AAPGBulletin, v. 78, p. 97–115.

Cade, C.A., I.J. Evans, and S.L. Bryant, 1994, Analysisof permeability controls—a new approach: ClayMinerals, v. 29, p. 491–501.

Cochran, J.R., 1983, A model for the development ofthe Red Sea: AAPG Bulletin, v. 67, p. 41–69.

Cole, G.A., M.A. Abu-Ali, H.I. Halpern, W.J. Carrigan,R. Savage, R.J. Scolaco, and S.H. Al-Sharidi, 1995, Thesource rock geochemistry of the Midyan and Jaizanbasins of the Red Sea, Saudi Arabia: Monam, Bahrain,Gulf Petrolink Selected Papers from the Middle EastGeoscience Conference, Bahrain, April 25–27, 1994, p. 307–319.

Coleman, R.G., 1993, Geological evolution of the RedSea: Oxford Monographs on Geology and Geo-physics, No. 24: New York, Oxford University Press,186 p.

Crossley, R., C. Watkins, M. Raven, D. Cripps, A. Car-nell, and D. Williams, 1992, The sedimentary evolu-tion of the Red Sea and Gulf of Aden: Journal ofPetroleum Geology, v. 15, p. 157–172.

Davidson, I., S. Al-Kadasi, A.K. Al-Subbary, J. Baker, S.Blakey, D. Bosence, C. Dart, R. Heaton, K. McClay, M.Menzies, G. Nichols, L. Owen, and A. Yelland, 1994,Geological evolution of the southeastern Red Sea riftmargin, Republic of Yemen: Geological Society ofAmerica Bulletin, v. 106, p. 1474–1493.

Davidson, I., D. Bosence, I. Alsop, and M.H. Al-Aawah, 1995, Deformation and sedimentationaround active Miocene salt diapirs on the TihamaPlain, NW Yemen, in G.I. Alsop, D.J. Blundell, andI. Davidson, eds., Salt tectonics: Geological SocietySpecial Publication 100, p. 1–18.

Davison, A., and D.C. Rex, 1980, Age of volcanismand rifting in southwest Ethiopia: Nature, v. 283,p. 657–658.

Egloff, F., R. Rihm, J. Makris, Y.A. Izzeldin, M. Bob-sien, K. Meier, P. Junge, T. Norman, and W. Warsi,1991, Contrasting structural styles of the easternand western margins of the southern Red Sea: the1988 SONNE experiment: Tectonophysics, v. 198, p. 329–353.

Ehrenberg, S.N., 1990, Relationship between diagenesisand reservoir quality in sandstones of the Garn For-mation, Haltenbanken, mid-Norwegian continentalshelf: AAPG Bulletin, v. 74, p. 1538–1558.

El-Anbaawy, M.I.H., M.A.H. Al-Anwah, K.A. Al-Thour, and M.E. Tucker, 1992, Miocene evaporites of

Poroperm Prediction for Wildcat Exploration Prospects: Miocene Epoch, Southern Red Sea 175

Porosity (%)

0.1

1

10

100

1000

10000

100000

0 10 20 30 40

Per

mea

bili

ty (

md

)

C

C

F

M

P

Figure 13. Poroperm data for halite cemented andcleaned sandstones from Antufash-1. C = com-paction/quartz cementation curve for fine (200 mm),moderately sorted sandstone. F = poroperm relation-ship for pore-filling clay cemented sandstone withabout 10% pore-filling clay. Porosity of 20% derivedfrom Figure 8 (1800 m burial). M = measuredporoperm of the Antufash-1 sandstones after halitedissolution; P = predicted permeability.

Page 189: Reservoir Quality Prediction in Sand and Carbonates

176 Gluyas and Witton

the Red Sea rift, Yemen Republic: sedimentology ofthe Saly Halite: Sedimentary Geology, v. 81, p. 61–71.

Evans, J., S.L. Bryant, and C.A. Cade, this volume,Modeling the effects of diagenetic cements onsandstone permeability, in J.A. Kupecz, J. Gluyas,and S. Bloch, eds., Reservoir quality prediction insandstones and carbonates: AAPG Memoir 69, p. 91–102.

Girdler, R.W., and P. Styles, 1978, Seafloor spreadingin the western Gulf of Aden: Nature, v. 271, p. 615.

Gluyas, J.G., 1985, Reduction and prediction of sand-stone reservoir potential, Jurassic North Sea: Philo-sophical Transactions of the Royal Society, v. A315,p. 187–202.

Gluyas, J.G., and C.A. Cade, this volume, Sand com-paction, in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones and car-bonates: AAPG Memoir 69, p. 19–28.

Gluyas, J.G., and M.L. Coleman, 1992, Material flux andporosity changes during diagenesis: Nature, v. 356,p. 52–53.

Gluyas, J.G., and A.J. Leonard, 1995, Diagenesis ofthe Rotliegend Sandstone: the answer ain’t blowin’in the wind: Marine and Petroleum Geology, v. 12,p. 491–497.

Gluyas, J.G., and N.H. Oxtoby, 1995, Diagenesis: ashort (2 million year) story—Miocene sandstones ofCentral Sumatra, Indonesia: Journal of SedimentaryResearch, v. A65, p. 513–521.

Gluyas, J.G., A.G. Robinson, D. Emery, S.M. Grant, andN.H. Oxtoby, 1993, The link between petroleumemplacement and sandstone cementation: GeologicalSociety of London, Petroleum Geology of NWEurope, Proceedings of the 4th Conference, March29–April 1, 1992, Barbican, London, p. 1395–1402.

Hall, S.A., 1989, Magnetic evidence for the nature ofthe crust beneath the southern Red Sea: Journal ofGeophysical Research, v. 94.

Heaton, R.C., M.P.A. Jackson, M. Bamahmoud, andA.S.O. Nani, 1993, Superimposed Neogene extension,contraction and salt canopy emplacement in theYemeni Red Sea (abs.): AAPG Hedberg ResearchConference on Salt Tectonics, September 13–17, Bath,England, Abstract Volume.

Hughes, G.W., and Z.R. Beydoun, 1992, The Red Sea—Gulf of Aden: biostratigraphy, lithostratigraphyand palaeoenvironments: Journal of PetroleumGeology, v. 15, p. 135–156.

Hughes, G.W., O. Varol, and Z. Beydoun, 1991, Evi-dence for Middle Eocene rifting of the Gulf of Aden

and Late Oligocene rifting in the southern Red Sea:Marine and Petroleum Geology, v. 8, p. 354–358.

Kazmin, V., 1973, Geological map of Ethiopia, AdisAbaba: Ministry of Mines, Geological Survey ofEthiopia, scale 1:2,000,000.

Loucks, R.G., M.M. Dodge, and W.E. Galloway, 1984,Regional controls on diagenesis and reservoir qual-ity in Lower Tertiary sandstones along the TexasGulf Coast, in D.A. McDonald and R.C. Surdam,eds., Clastic diagenesis: AAPG Memoir 37, p. 15–45.

Merla, G., 1979, A geological map of Ethiopia and Soma-lia: Rome, Consiglio Naz. Reirche, scale 1:2,000,000.

Mitchell, D.J.W., R.B. Allen, W. Salana, and A.Abouzahu, 1992, Tectonostratigraphic frameworkand hydrocarbon potential of the Red Sea: Journalof Petroleum Geology, v. 15, p. 187–210.

Primmer, T.J., 1993, Regional diagenesis in sandstonesand controls on reservoir quality: a petroleumindustry perspective (abs.): Abstracts of the 5thCambridge Clay Diagenesis Meeting of the ClayMinerals Group, March 25–26, 1993.

Primmer, T.J., C.A. Cade, I.J. Evans, J.G. Gluyas, M.S.Hopkins, N.H. Oxtoby, P.C. Smalley, E.A. Warren,and R.H. Worden, this volume, Global patterns insandstone diagenesis: their application to reservoirquality prediction for petroleum exploration, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 61–78.

Robinson, A.G., and J.G. Gluyas, 1992, Model calcula-tions of sandstone porosity loss due to compactionand quartz cementation: Marine and PetroleumGeology, v. 9, p. 319–323.

Rønnevik, H., S. Eggen, and J. Vollset, 1983, Explorationof the Norwegian Shelf, in J. Brooks, ed., Petroleumgeochemistry and exploration of Europe: GeologicalSociety of London Special Publication 12, p. 71–94.

Schmoker, J.W., and D.L. Gautier, 1988, Sandstoneporosity as a function of thermal maturity: Geology,v. 16, p. 1007–1010.

Small, J.S., D.L. Hamilton, and S. Habesch, 1992, Experi-mental simulation of clay precipitation within reser-voir sandstone 1: techniques and examples: Journal ofSedimentary Petrology, v. 62, p. 508–519.

Sudan, 1963, Sudan geological map, Khartoum: SudanSurvey Department, 3d edition, scale 1:4,000,000.

U.S. Geological Society, 1963, Geologic map of theArabian Peninsula: U.S. Geological Society Misb.Geol. Inves. Map 1-270-A Washington, D.C.: U.S.Geological Survey, scale 1:2,000,000.

Page 190: Reservoir Quality Prediction in Sand and Carbonates

177

Chapter 12

Porosity–Depth Trends in Deeply BuriedUpper Jurassic Reservoirs in the Norwegian

Central Graben: An Example of PorosityPreservation Beneath the Normal Economic

Basement by Grain-Coating MicroquartzMogens Ramm1

Norsk Hydro Research CentreBergen, Norway

Arne W. ForsbergNorsk Hydro Exploration

Stabekk, Norway

Jens S. JahrenDepartment of Geology, University of Oslo

Oslo, Norway

ABSTRACT

Successful prior-to-drilling prediction of anomalously good reservoir quali-ty in prospects at deep burial requires an understanding of diageneticprocesses and quantitative models on how porosity is related to sandstonecomposition and to burial history. Quartz cementation and compaction are,in many cases, the most important porosity-reducing processes in quartz-and feldspar-rich arenites, capable of destroying all useful porosity duringburial toward 4000 m. Hence, the recognition of factors that may hinderporosity loss by these processes, and thereby preserve good reservoir qualityto depths beneath those usually considered as economic basement, is crucialduring prospect evaluation of deep structures.

In two deep (>4000 m) oil discoveries in Upper Jurassic sandstones in theNorwegian Central Graben, high porosity (>20%) appears to be preserveddue to the presence of a ubiquitous microquartz coating on frameworkgrains, and not due to any burial history-dependent factor such as highpore pressure, low thermal maturity, or early oil emplacement. In thesesandstones, the microquartz coating has hindered quartz precipitation andlate diagenetic chemical compaction. In interbedded sandstones without

1 Present address: Norsk Hydro Exploration, Stabekk, Norway.

Ramm, M., A.W. Forsberg, and J.S. Jahren, 1997,Porosity–depth trends in deeply buried UpperJurassic reservoirs in the Norwegian CentralGraben: an example of porosity preservationbeneath the normal economic basement bygrain-coating microquartz, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality pre-diction in sandstones and carbonates: AAPGMemoir 69, p. 177–199.

Page 191: Reservoir Quality Prediction in Sand and Carbonates

178 Ramm et al.

INTRODUCTION

In the Norwegian sector of the North Sea and in thehydrocarbon provinces off mid-Norway, the porosityin most arenitic sandstone reservoirs appears to followlinear porosity vs. depth trends that can be expressedas φ= φ0 – G ×Z, where φ0 is the porosity at the time ofdeposition, G is the porosity depth gradient, and Z isburial depth (Selley, 1978; Bjørlykke et al., 1986, 1989;Ehrenberg, 1990; Ramm and Bjørlykke, 1994). Further-more, it appears that much of the deviation from suchtrends is related to variations in petrographic charac-teristics, such as variations in total clay content (Rammand Bjørlykke, 1994) or the presence or absence of claycoatings (Ehrenberg, 1993). However, empirical datafrom the Norwegian shelf also document that devia-tions from the general trends may be related to burialhistory-dependent factors; for example, due toretarded compaction in highly overpressured reser-voirs (Ramm, 1992; Ramm and Bjørlykke, 1994) ordue to variations in degree of quartz cementation dueto differences in the time-temperature–dependent“diagenetic maturity” (Walderhaug, 1994a).

During the evaluation of deeply buried (4000–5000 m)structures in the Cod Terrace area, in the NorwegianCentral Graben (Figure 1), it was observed that thereservoir quality in Upper Jurassic sandstones fromthe Mime field and neighboring structures showedsome large, unpredicted porosity variations. Of specialinterest in this respect was the observation that somesandstone intervals have porosities significantlyhigher than usually found at depths of 4000 m in theNorth Sea. Most important, Jurassic reservoir sand-stones from the Norwegian shelf follow approxi-mately the same porosity vs. depth trend (Figure 2A).However, only some of the Upper Jurassic sandstonesfrom the Cod Terrace area follow this trend, and it isapparent that one group of sandstones, buried todepths of ~4000 m, has significantly better porositythan the “normal trend.” These high-porosity outliersare replotted with other high-porosity outliers from

the Norwegian shelf in Figure 2B. In these other high-porosity outliers, the high porosity can either beexplained with retarded compaction due to high porepressures or with retarded quartz cementation due tochlorite coatings on the detrital grains. Apparently, theUpper Jurassic sandstones have porosities of the samemagnitude as the other high-porosity outliers, butthese sandstones do not contain chlorite coatings, andthe reservoirs are not highly overpressured. Hence,other explanations must be sought, and a successfulprior-to-drilling porosity prediction in this arearequires a good quantitative model on how porosity isrelated to sandstone composition and burial history; italso requires recognition of the factors or processesthat preserve good reservoir quality to depths beneaththose usually considered as the economic basement.

REGIONAL SETTING

The Norwegian part of the Central Graben area inthe southernmost parts of the Norwegian shelf repre-sents one of the richest hydrocarbon provinces in theNorth Sea; hydrocarbon discoveries have been madeboth within Cretaceous chalk reservoirs and withinUpper Jurassic and Triassic sandstone reservoirs. Theearly exploration in the area was concentrated on Cre-taceous chalk plays; the important discoveries Cod,Ekofisk, Eldfisk, Edda, Tommeliten, and Tor fieldswere made during the first decade of hydrocarbonexploration in the Norwegian North Sea. The UpperJurassic sandstone plays, however, have a shorter his-tory. The first well in block 7/12 tested a Cretaceouschalk play and was abandoned about a hundredmeters above the subsequently discovered UpperJurassic Ula field. The chalk interval was dry, however,and the block was relinquished. A renewed interest inthe area during the late 1970s and early 1980s led to theUla discovery (70 ×106 Sm3 recoverable oil) in 1976 (1 Sm3 = standard cubic meter = 6.3 barrel), the Gydadiscovery (32 ×106 Sm3 recoverable oil) in 1980, and the

microquartz coating, the porosity is low (<10%) due to extensive quartzcementation. The microquartz coating appears within specific isochronouslayers, and its presence is probably caused by input of amorphous silica(volcanic glass and sponge spicules) during deposition. The recognition ofthe inhibiting effect of this coating on quartz cementation, combined withquantitative models on the relationship between sandstone compositionand diagenetic processes such as compaction and quartz cementation,allows confident porosity predictions. Hence, future porosity prediction indeeply buried Upper Jurassic sandstone in this area should focus on estab-lishing sedimentological models addressing prediction of sandstone facieswithin intervals deposited during periods with high amorphous silica production and deposition.

Page 192: Reservoir Quality Prediction in Sand and Carbonates

minor Mime discovery (0.6 ×106 Sm3 recoverable oil) in1982. These three reservoirs have been developed forproduction; the Ula field came onstream in 1986, theGyda Field in 1990, and the Mime Field in 1992. Today,these reservoirs represent the deepest that have beenbrought into production in the Norwegian sector of theNorth Sea (Figure 3) and, as such, the good reservoirproperties, particularly in the two deepest reservoirs(the Mime and Gyda fields), are very noteworthy.

GEOLOGICAL FRAMEWORK

The Upper Jurassic section in the Central Graben wasdeposited in an open shelf environment following anOxfordian relative sea level rise (Home, 1987; Forsberg

et al., 1994; Oxtoby et al., 1995). Forsberg et al. (1994)introduced a sequence stratigraphical nomenclature forthe Upper Jurassic sections in the area based on a multi-disciplinary genetic sequence stratigraphical approach.They were able to draw timelines through the shallowmarine sandstone units from the Cod Terrace area (for-merly termed the Ula Formation) and deeper marinemuddy sequences (formerly termed the Haugesundand Farsund formations) of the central parts of the FedaGraben. Accordingly, five Upper Jurassic sequenceswere identified (Figure 4).

The Cod Terrace area was subaerially exposedbefore the Oxfordian transgression during the Earlyand Middle Jurassic, and the Upper Jurassic strata arefrequently overlying Triassic rocks. Only locally are

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 179

Figure 1. Location mapshowing depth (two-waytraveltime in milliseconds)to the base-Cretaceousunconformity, Cod Terracearea, Norwegian CentralGraben.

Page 193: Reservoir Quality Prediction in Sand and Carbonates

180 Ramm et al.

Figure 2. Porosity vs. depth plots showing the 75% value of He-porosity measurements within different Jurassicsandstone units from wells from the North Sea and the Haltenbanken areas, offshore Norway. (A) Upper Jurassicsandstones from the Cod Terrace area compared to normal-porosity reservoir sandstones. The “normal” porositydepth trend is expressed as φ= 47.4–0.0089 ×Z, r2 = 0.87, N = 214. Only clean arenitic sandstone units not containingchlorite coatings and not representing highly overpressured reservoirs (pressure gradient >18 MPa/km) are shown.(B) Upper Jurassic high-porosity sandstones from the Cod Terrace area compared to high-porosity outliers fromother areas on the Norwegian Shelf.

Page 194: Reservoir Quality Prediction in Sand and Carbonates

fluvial deposits of the Bathonian to Bajocian BryneFormation observed. The reservoir sandstones in theCod Terrace area accumulated after the Late Oxfor-dian transgression, mainly in topographic depressionson the downfaulted side of the major Ula-Gyda faultzone (Home, 1987), which separates the SørvestlandetHigh from the Cod Terrace (Figure 1).

The Upper Jurassic sequences comprise mudstones tofine- and medium-grained sandstones, deposited in anoffshore marine shelf setting. Most sandstones containappreciable fine-grained clay material; primary sedi-mentary structures are rare due to extensive bioturba-tion. The grain size is rather uniform, varying mainlybetween 100 and 300 µm. Textural cleaning upwardtrends are pronounced and formed as response to rela-tive sea level changes. Abrupt flooding events followedby slow sediment aggradation, with the most wellsorted clean sandstones deposited on top of eachsequence, are well illustrated by the gamma-ray logs. Alithostratigraphic subdivision of the Upper Jurassic sec-tion, based on the sequence stratigraphical nomencla-ture of Forsberg et al. (1994), is shown in Figure 5, and amore detailed definition of the units is presented inAppendix A.

POROSITY TRENDS

Porosity vs. Depth

High- and low-porosity zones are stratigraphicallycorrelatable in wells from the Gyda area (Figure 5A).High-quality reservoir zones are seen in units C12 andB2, whereas nonporous sandstones are seen in unitsC14, C11, B1, and A (Figure 5A). The transgressive

sandstone in unit A1, cored in well 2/1-2, has moderateporosity compared to its burial depth (3340 m RKB [rel-ative to Kelly Bushing]). Low porosities in the muddyintervals of units C14 and B1 are probably related to thehigh clay content (e. g., high gamma-ray signals) ofthese rocks. The low porosity in the cleaner interval ofunit C11, however, contrasts the good porosity in theapparently analogous sandstones of unit C12 and theupper parts of unit B.

The porosity is more uniformly distributedthroughout the cored sections in the sandstones fromthe Ula area (Figure 5B). Low porosities are observedwithin muddy intervals and in carbonate cementedzones. The wells from the central part of the Ula field(i.e., wells 7/12-2, 7/12-4, and 7/12-6) have porositiesof ~20% at ~3400 m RKB. The more deeply buried sec-tions of well 7/12-7 and 7/12-5, however, have porosi-ties of only 15% and 12% at 3800 m and 3900 m,respectively. Wells 7/11-5 and 7/11-6 also show largeporosity variations within apparently homogeneoussandstones. A reversed relationship between porosityand clay content is indicated in units A and B in well7/11-5. Here the porosity is higher in unit B than inunit A, although the gamma-ray log signals indicatecleaner sandstones in upper parts of A than in B.

The lithostratigraphical units are divided into three“reservoir-quality facies” with respect to the relationbetween porosity and depth. Facies 1, the high-porosityoutliers, is characterized by high porosity (>20%) atdepth >4 km and represents units B2 and C12. Facies 2,the normal-porosity sandstones, are those characterizedby low gamma-ray signals and low porosity at greatburial depth (>4000 m) (i.e., units A1, A3, and C11). Facies3, the poor reservoir quality sandstones, are those with

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 181

Figure 3. Burial depths for oil and gas fields off Norway.

Page 195: Reservoir Quality Prediction in Sand and Carbonates

182 Ramm et al.

high gamma-ray signals and poor reservoir quality at alldepths (Table 1). Least-squares regression lines based onthe porosity vs. depth relationship for the three reser-voir-quality facies are shown in Figure 6. Near 3500 m,the porosity is approximately the same in facies 1 and 2,where the porosity in both clean, low gamma-ray sand-stone facies is ~25%. Below 3500 m, however, the poros-ity in the two facies differentiates. In the facies 2sandstones, the porosity–depth gradient is steep, and at4500 m these sandstones have porosities <10%. To thecontrary, the facies 1 sandstones lose porosity more gen-tly with depth, and porosities near 20% are preserved to4500 m. The muddy facies 3 sandstones have porosities<15% at all depths, and these units are in essenceunprospective at depths below 3000 m.

Porosity vs. Pore Fluid Composition

The porosities of facies 1 and 2 sandstone unitsfrom oil zones and from the water legs/dry wells areplotted vs. depth in Figure 7. The figure depicts no sig-nificant and systematic difference in porosity betweenwater- and hydrocarbon-saturated reservoirs.

The porosity difference across the oil-water contactis particularly large in well 7/11-5. In the oil-filled unitB and the water-filled unit A, the 75% porosities are23.6% and 12.3%, respectively. In well 1/3-3, however,

the oil-water contact is located about 10 m below theboundary between the low-porosity unit C13 and thehigh-porosity unit C12. However, a minor decrease inporosity is observed across the oil-water contact at 4221m RKB. The average porosity between 4213–4221 m is23.7% (standard deviation 2.0%), whereas between4221 and 4247.5, it is 22.1% (standard deviation2.3%). The standard deviations in the two subpopu-lations are approximately equal, and a simple one-way analysis of variance can be used to reject, at the95% confidence level, the hypothesis that the poros-ity difference above and below the water zone (1.6%)is not statistically different (the calculated φvalue forthe data is 9.6, with 1 and 102 degrees of freedom,which is larger than φ(0.05;1102) = 3.9). However, thisdoes not necessarily mean that it is the difference inthe pore-fluid composition that causes the porositydifference.

Porosity Related to Lithology

The high-porosity zones in units C12 and B2 in well2/1-6 contain extensively bioturbated, fine-grainedsandstones. The low-porosity unit C14 comprisesmuddy graywackes and siltstones, whereas the low-porosity sandstone in unit C11 is a relatively clean,medium-grained sandstone. The coarsening/cleaning-upward sequence in unit B indicates a clear inverse

Figure 4. Upper Jurassic stratigraphy in the Cod Terrace area, showing the chronostratigraphic andsequence stratigraphic division of the Upper Jurassic strata and distribution of shales and sandstonesin the Feda Graben and Cod Terrace area (modified from Forsberg et al., 1994). Letters refer tosequences discussed in the text.

Page 196: Reservoir Quality Prediction in Sand and Carbonates

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 183

Figu

re 5

. Wel

l cor

rela

tion

bet

wee

n w

ells

fro

m th

e G

yda

area

(A) a

nd

Mim

e-U

la a

rea

(B) s

how

ing

gam

ma-

ray

log

and

cor

e-p

oros

ity

pan

els

for

the

ind

ivid

ual

wel

ls. T

he

ran

ge in

the

por

osit

y p

anel

s is

0%

–30%

. Loc

atio

n o

f w

ells

is s

how

n in

Fig

ure

1.

Page 197: Reservoir Quality Prediction in Sand and Carbonates

184 Ramm et al.

correlation between clay content and porosity. How-ever, this correlation is disrupted by the clean, low-porosity sandstones in unit C11.

Inverse correlation between clay content and poros-ity is also observed in units B and C11–3 of well 7/11-5.The gamma-ray logs from this well indicate a slightlyhigher clay content in the low-porosity unit C11–3 than

in the high-porosity unit B (Figure 5). The upper partof unit A, however, resembles unit C11 in well 2/1-6and has very low porosity, in spite of a very low claycontent, as indicated by the gamma-ray signal.

In unit A in well 7/12-4, a strong negative correla-tion between clay content and porosity is indicatedand, as such, the unit resembles unit B of wells 7/11-5

Table 1. Characteristics of Wells.*

Reservoir Hydro- Depth Mean BurialQuality carbon Interval Thickness Depth 75-Percentile Number

Well Unit Facies Present (m RKB***) (m) (m RSF***) Porosity of Plugs

1/3-3 C14 3 4138–4148 10 4049 8.2 24C13 2 + 4181–4200 19 4098 14.6 61C12 1 +/- 4210–4248 38 4135 23.8 115

2/1-2 A2 3 3318–3329 11 3230 9.0 33A1 2 - 3330–3336 9 3239 14.8 19

2/1-3 C11 2 + 3823–3832 9 3731 15.3 24B2 1 + 3840–3862 22 3755 23.7 53B1 3 3880–3888 8 3788 5.4 18

2/1-4 C11 2 + 4036–4090 54 3972 13.2 163B2 1 + 4095–4125 30 4021 18.4 79B1 3 4132–4138 6 4044 5.9 20

2/1-6 C14 3 4171–4200 29 4094 8.9 65C12 1 - 4202–4245 43 4133 21.6 126C11 2 - 4250–4315 65 4201 9.2 49B2 1 - 4320–4350 30 4244 16.7 90B1 3 4361–4376 15 4277 8.5 16

2/1-8 B2 1 + 3898–3923 25 3807 19.8 60B1 3 3931–3955 24 3840 4.1 63A 3 3955–3981 26 3865 3.4 65

7/8-3 Ula Fm** (2) + 3731–3768 37 3618 13.7 1137/11-5 C11-3 2 + 4159–4171 22 4060 16.9 26

B 1 + 4171–4191 20 4083 23.6 67A 2 - 4205–4238 33 4117 12.3 112

7/11-7 C11-3 2 - 4100–4111 11 4005 13.5 46B 1 - 4110–4131 21 4020 17.8 74A 2 - 4131–4145 14 4037 11.7 55

7/11-7 C2 (1) - 4549–4557 8 4435 18.0 25C2 (2) - 4558–4565 7 4443 8.6 15

7/11-9 A? (2) - 4172–4177 5 4068 13.9 167/12-2 C11-3 2 + 3385–3410 25 3302 21.6 63

B 1 + 3410–3476 66 3347 23.7 1387/12-4 B 1 + 3450–3492 42 3375 20.5 102

A3 2 + 3492–3510 18 3406 19.0 48A2 3 3511–3525 14 3423 12.5 30

7/12-5 A 2 +/- 3850–3900 50 3772 12.2 1397/12-6 C11-3 2 + 3407–3434 27 3327 20.9 62

B 1 + 3434–3474 40 3361 21.0 106A3 2 + 3474–3507 33 3397 22.0 93

712-7 A-B 2 +/- 3800–3842 42 3737 16.8 11223/27-3 Ula Fm (1) + 4010–4047 37 3920 22.1 9623/27-4 Ula Fm (1) - 3405–3425 20 3309 24.2 6323/27-6 Ula Fm (1) - 3869–3909 40 3778 25.2 126

*Measured depth, burial depth, presence of hydrocarbons, and porosity of individual sandstone units.**Ula Fm = Ula Formation.***m RKB = depth relative to Kelly Bushing; m RSF = depth relative to sea floor.

Page 198: Reservoir Quality Prediction in Sand and Carbonates

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 185

Figure 6. Porosity in individual sandstone units vs. depth. The sandstone units are divided into threereservoir-quality facies showing different porosity depth relationships (Table 1). Facies 1 (high-porosity sandstones): φ= 45 ×e(–0.196 ×Z/1000); Facies 2 (normal-porosity sandstones): φ= 47.3-0.0085 × Z;Facies 3 (poor reservoir quality mudstone): φ= 45 ×e(–0.490 ×Z/1000). The exponential regression lines forthe high- and poor-porosity facies are obtained by locking the pre-exponential factor to 45%.

Figure 7. Porosity in facies 1 (high-porosity) sandstones and 2 (normal-porosity) sandstones replot-ted with respect to the presence or absence of hydrocarbons.

Page 199: Reservoir Quality Prediction in Sand and Carbonates

186 Ramm et al.

and 2/1-6. Within unit B, however, no such clear corre-lation between porosity and clay content is indicated.Except for the low-porosity levels corresponding tocarbonate cemented layers and the reduced porosity inthe lower part of unit A, the porosity is rather homoge-neously distributed throughout the cored section inthis well.

Porosity vs. Bulk Mineralogy

Eighty-eight samples from wells 7/11-5, 7/11-6,7/12-4, and 2/1-6 have been analyzed for their bulkmineralogy composition by quantitative XRD (X-ray diffraction) measurements (methods used are docu-mented by Ramm, 1991). To test the correlation betweenbulk mineralogy and porosity, correlation coefficientsbetween porosity and mineral content from the com-plete XRD data set and a number of sample subsetswere examined (Table 2).

When all samples were considered (Table 2, column1), there is a significant positive correlation betweenporosity, quartz, and feldspar content, but there is a neg-ative correlation between porosity, clay, and calcite con-tent. The positive correlation between quartz contentand porosity is influenced by the low quartz contents inseven carbonate cemented samples and in four mud-stones samples, all having low porosity.

When the mudstones and the carbonate cementedsamples (>10% total carbonate) are excluded, porosityis negatively correlated to depth and clay content butpositively correlated with feldspar content. In thissample subset, there is no positive correlation betweenquartz content and porosity (Table 2, column 2).

The correlation between porosity and depth ismainly caused by differences in porosity betweensamples from well 7/12-4 (3450–3525 m) and the sam-ples from the three other wells (4100–4355 m). Hence,considering the data from the three deeper wells givesa possibility to assess the mineralogical influence onthe porosity variations at deep burial (Table 2, column3). Although significant negative correlation betweenporosity, clay, and calcite content, and positive corre-lations between porosity and feldspar content, areindicated, the correlation coefficients are low, and fewstatistically significant values are found.

Figure 8 depicts the variations in the Clay Index(the ratio of total clays to quartz plus feldspar con-tent) and quartz content vs. porosity. It is observedthat most samples follow a trend of reducing porositywith increasing Clay Index. Some samples with littleclay have very low porosity, however. All of thesesamples are extensively cemented either with carbon-ate or quartz cement. When these samples areexcluded, systematic and highly significant relation-ships between porosity bulk mineralogy areobserved (Table 2, column 4). Significant and nega-tive correlation between porosity and clay content isparticularly apparent.

The quartz cemented intervals in unit C11 in well2/1-6, and upper parts of unit A in well 7/11-5 do notfollow the same trend between porosity and clay

content as do the other samples. The samples from thelow gamma-ray interval 4201–4345 m in well 2/1-6,representing units B2, C11, and C12, show distinctly dif-ferent relations between porosity and bulk mineralogy(Table 2, column 5). These samples all contain little claymaterial, but the porosity is very variable. Within thisgroup of samples, the quartz content shows a strong neg-ative correlation with porosity. This relation is illustratedin Figure 8B, where the quartz cemented samples fromunit C11 in well 2/1-6 and those from the upper part ofunit A in well 7/11-5 cluster in the lower right corner ofthe diagram and are characterized by having low poros-ity and very high contents of quartz plus feldspar.

Petrographic Observations from Thin Sections

Thin sections from wells 7/11-5, 7/11-6, 7/11-10S,7/12-4, and 2/1-6 have been studied and point counted.The point counting was done with emphasis on esti-mating the amount of intergranular (primary) andintragranular (secondary) porosity and the amount ofintergranular cements (mainly quartz and carbonate).The petrographical characteristics of the different strati-graphical units are indicated through a brief descriptionof samples from well 2/1-6 in Appendix B.

Intergranular volume (IGV) vs. cement diagramsincluding data from well 2/1-6 are shown in Figure 9.The compactional and cementational porosity loss(COPL and CEPL, respectively) are estimated by equa-tions 1 and 2, which are modified versions of thosepresented by Ehrenberg (1989)

(1)

(2)

where φ0 is the original porosity (which here isassumed to be 45%), TC is the total cement, and IP isthe intergranular porosity. All parameters areexpressed as volume percentages of total rock volume.

Most samples (e.g., those from units B, C12, andC14) have little quartz cement and have lost most oftheir intergranular porosity by compaction. Theirpresent intergranular porosity is inversely correlatedwith the matrix content. Substantial amounts ofquartz cement are observed in the samples from unitC11. The samples from well 2/1-6 may be dividedinto three groups. The well-sorted arenites from unitsC12 and B2 are characterized by low contents ofmatrix and quartz cement and by high porosity.These sandstones contain 25%–30% intergranularporosity plus cement and ~7% cement. According toequations 1 and 2, they have lost ~21%–27% porosity(~50% of the original porosity) by compaction and5%–6% (~10% of the original porosity) by cementa-tion. The fine-grained graywackes from units C14 andB1 are characterized by low porosity and low con-tents of quartz cement. On the average, they contain17% intergranular porosity plus cement and 5%

CEPL COPLTC

IP TC= ( ) ×

+φ0 –

COPLIP TC

IP TC=

( ) × +( )+( )φ

φ0

0––

–100

100

Page 200: Reservoir Quality Prediction in Sand and Carbonates

cement and have, according to equations 1 and 2, lost~34% porosity (75% of their original porosity) bycompaction and 3% (~7% of the original) by cementa-tion. The quartz cemented arenites from unit C11 arecharacterized by low porosity and matrix content andhigh content of quartz cement. On the average, theycontain ~20% intergranular porosity and 15% cementand have, according to equations 1 and 2, lost ~30%(67% of the original) porosity by compaction and 10%(25% of the original) by cementation.

Petrographic Observations Using Scanning Electron Microscopy

The petrographic observations from bulk mineral-ogy analyses by XRD and from thin sections revealedthat much porosity variation can be related to varia-tions in the clay content. However, the samples fromunit C11 in well 2/1-6 and the uppermost part of unitA in well 7/11-5 have low porosity in spite of low claycontent, and this is due to the extensive quartz cemen-tation. Sample chips from wells 2/1-6 and 7/11-5 havebeen examined in SEM in order to describe texturesthat might explain why some of the clean samples areextensively quartz cemented, whereas others are not.Secondary electron images of characteristic samplesfrom high-porosity zones in well 2/1-6 are presentedin Figure 10. The samples are characterized by littlepore-occluding cement and high preserved primaryintergranular porosity. It is observed, however, that

clean quartz grain surfaces are not present. All grainsare coated with clay minerals, and more commonly bysmall (1 mm) microquartz crystals. Occasionally 10- to50-mm-large euhedral overgrowths stand out from thecoated framework grains.

Larger euhedral quartz overgrowths are extremelyrare in the high-porosity samples, and it appears thatthere is a close relationship between the occurrence ofthe microquartz coating and amount of euhedral quartzcement. Hence, minor porosity loss caused by quartzcementation may be due to efficient inhibition of latediagenetic growth of quartz cement by the clay andmicrocrystalline quartz coating on framework grains.Thus, the high porosity in the clean sandstones of unit Bin well 7/11-5 and units C12 and B2 of well 2/1-6appears to be related to inhibited quartz cementation bythe coating. In unit C11 of well 2/1-6 and the upper partof unit A in well 7/11-5, quartz precipitation has notbeen inhibited; much porosity is destroyed by chemicalcompaction and quartz cementation in these sandstones.

Constraints on the Quartz Precipitation from FluidInclusion Homogenization Temperatures

Quartz cement is the volumetrically most importantcement in the deeply buried, clean sandstones with lowporosity, but nearly absent in the good-porosity sand-stones. Quartz cement accounts for ~15% in unit C11 inwell 2/1-6, 20% in unit A in well 7/11-5 (Walderhaug,1994b), and 5% in well 7/12-6 (Nedkvitne et al., 1993).

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 187

Table 2. Correlation Coefficients Between Porosity and Mineral Content.

All Samples–– Quartz-Cemented Only IntervalMudstones Only Samples from 4201-4345

and Carbonate Wells 2/1-6 + 7/11-5 + 2/1-6 mKB inAll Samples Cemented 7/11-5 + 7/11-6 Excluded Well 2/1-6

(n = 88) (n = 77) (n = 51) (n = 41) (n = 11)

Depth –0.3031* –0.3914† –0.2069 –0.0018 –0.4115Chlorite/Quartz –0.3504† –0.3668* –0.3041‡ –0.5321* –0.1920Chlorite –0.3787† –0.3819† –0.3107‡ –0.5521† –0.2095Illite/Quartz –0.3813† –0.3143* –0.1950 –0.4826* 0.2682Illite –0.3881† –0.3394* –0.2027 –0.5274† 0.2572Clay Index –0.3894† –0.4891† –0.3925‡ –0.8490† 0.3224Quartz 0.3259 –0.1650 –0.1524 0.0765 –0.6529‡

K-Feldspar/Quartz 0.0835 0.3380‡ 0.2458‡ 0.2795‡ 0.6568‡

K-Feldspar 0.3951† 0.3361* 0.2135 0.3980 0.1021Albite/Quartz 0.0121 0.2525‡ 0.2645‡ 0.0944 0.4392Albite 0.4083† 0.3234* 0.3717* 0.2849‡ 0.3306Calcite/Quartz –0.4371† –0.1327 –0.2808‡ –0.2733‡ –0.2900Calcite –0.4537† –0.1443 –0.3541 –0.2890‡ –0.5252‡

Ankerite/Quartz –0.0699 –0.1297 0.0434 –0.1410 0.6065‡

Ankerite –0.0160 –0.1665 0.0023 –0.1608 0.5382‡

Siderite/Quartz –0.1313 –0.0450 –0.0469 –0.0688 0.2321Siderite –0.0242 –0.0466 –0.0896 –0.0656 –0.1048Pyrite/Quartz –0.0785 –0.1195 –0.0361 –0.1655 –0.0937Pyrite –0.0392 –0.1284 –0.0467 –0.1894 –0.1689

† Significance level > 99.9%.* Significance level > 99%.‡ Significance level > 90%.

Page 201: Reservoir Quality Prediction in Sand and Carbonates

188 Ramm et al.

Figure 8. Bulk mineralogy vs. porosity. (A) Clay Index vs. He-porosity (Clay Index = sum of allphyllosilicates divided by the sum of quartz and feldspar, here the normalized ratio of the 19.8° 20XRD-peak to the normalized quartz and feldspar XRD-peaks) (B) Quartz content from bulk XRDvs. He-porosity. The regression line for the porosity vs. Clay Index trend (φ= 26.9 e(–3.46 x CI), r2 = 0.75)is based on samples without abundant quartz and/or carbonate cement.

Page 202: Reservoir Quality Prediction in Sand and Carbonates

The homogenization temperatures in petroleum inclu-sions in wells 7/12-6 and 7/11-5 are 30–50°C lowerthan in the aqueous inclusions (Figure 11), which prob-ably implies that the petroleum inclusions compriseundersaturated oils (with respect to gas), while theaqueous inclusions are saturated. The validity of fluidinclusion homogenization temperatures as a tool topinpoint the quartz precipitation temperatures has

been questioned by Osborn and Hazeldine (1993), whosuggested common stretching and resetting of thehomogenization temperatures toward equilibrationwith the bottom-hole temperature. A close correlationbetween maximum temperatures and homogenizationtemperatures may not necessarily reflect resetting, butmerely reflect the burial and temperature history ofsandstones as suggested by a model by Walderhaug

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 189

Figure 9. Diagrams modifiedfrom Houseknecht (1987) show-ing (A) the total cement vs. totalintergranular cement and poros-ity, and (B) quartz cement vs.intergranular porosity plusquartz cement in samples fromwell 2/1-6. These diagrams illus-trate the relative effect of com-paction and cementation onporosity. The most porous sand-stones in units C12 and B2 havelost about half their initialporosity by compaction, where-as the muddy sandstones fromunits C14 and B1 have lost morethan 75% of their initial porosityby compaction. Neither of thesesandstones have lost muchporosity by quartz cementation.The quartz-cemented samplesfrom unit C11 have also lost sig-nificantly more porosity by com-paction than by cementation.

Page 203: Reservoir Quality Prediction in Sand and Carbonates

190 Ramm et al.

Figure 10. Secondary electron images of rock chips from high-porosity sandstones in well 2/1-6. (A, C, and E)4209.00 m RKB; (B, D, and F) 4219 m RKB. Note absence of large euhedral quartz overgrowths and abundantmicrocrystalline quartz coatings on framework grains. Some larger (10–50 mm) overgrowths are observed.These larger crystals occur exclusively on quartz grains; where more than one occur on the same grain, they areparallel, indicating growth in optical continuity with the parent grain.

Page 204: Reservoir Quality Prediction in Sand and Carbonates

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 191

Figure 11. Fluid inclusion homogenization temperatures from quartz overgrowths. Data from Nedkvitneet al. (1993) and Walderhaug (1994b).

Page 205: Reservoir Quality Prediction in Sand and Carbonates

192 Ramm et al.

(1994a). According to this model, correlation betweenthe “lowest” homogenization temperature in a sand-stone and the present reservoir temperature may fol-low from the heating history rather than from resettingof the inclusions. If sandstones of similar age are buriedat different depths in neighboring structures, then thesandstones in the most deeply buried structure mayhave passed into deeper parts of the “quartz cementa-tion window” more rapidly than the sandstones in themore shallowly buried reservoir. Hence, the homoge-nization temperatures in the aqueous inclusions shownin Figure 11 may reflect the true trapping temperature,whereas a significant pressure correction is requiredfor the petroleum inclusions.

According to the above assumptions, it follows thatthe fluid inclusion homogenization temperatures fromwells 7/12-6 and 7/11-5 indicate that most of thequartz cement precipitation occurred at temperatures>120–130°C, which translates to burial depths >3000 m,and that precipitation has continued until the present(Figure 11). The overall high “lowest” homogenizationtemperatures in the two wells (and the higher “lowest”values in well 7/11-5 than in well 7/12-6) probablyreflect rapid and differential subsidence during theLate Oligocene to early Miocene. The area received~1000 m of overburden during this period, and theUpper Jurassic sandstones in wells 7/12-6 and 7/11-5were buried to ~2500 and 3000 m, respectively, duringthis short period. Hence, it is likely that little quartzprecipitated prior to or during this period of rapid bur-ial; most of the cement precipitated, at slightly differentburial depth and temperature in the two wells, after theperiod of rapid subsidence (Walderhaug, 1994a).

DISCUSSION

Burial Control on Porosity Variations

Theoretical models and empirical data from theNorwegian Shelf indicate that porosity–depth trendsare affected by pore pressure gradients and by time-temperature history (Ramm, 1992; Ramm and Bjør-lykke, 1994; Walderhaug, 1994a). Within the Jurassicsections on the Cod Terrace, the pore pressures aremoderately high and show overall small variations(Figure 12). Furthermore, sandstones within continu-ous sandstone intervals, representing identical pres-sure compartments, show large porosity variation.Hence, pore pressure variations may not explain thevarying porosity within the arenitic sandstone units.Similarly, differences in thermal maturation cannotexplain the large porosity variations observed withinsome of the continuous sandstone sequences.

Hydrocarbon emplacement has been claimed to haltor retard diagenetic processes and preserve porosityduring subsequent burial (Hancock and Taylor, 1978;Selley, 1978; Sommer, 1978; Gluyas et al., 1990, 1993).Many of the Upper Jurassic sandstone sequencesdescribed in this study contain hydrocarbon-saturatedpore spaces, and oil emplacement may accordingly besuggested to play a controlling role on the porosityvariations. The timing of the oil emplacement is critical

if an effect of such emplacement is expected. If the oilentered the sandstone after burial to approximately thepresent burial depth, porosity–depth trends would notbe expected. In the Ula field, the most significant periodof oil emplacement has been the last 3–5 m.y., andpetroleum migration is probably still continuing (Ned-kvitne et al., 1993). However, the Upper Jurassic sand-stones in well 7/11-5 have been buried to 700–800 m,and probably have been heated from <150°C to >160°Cduring the last 3–5 m.y.; the tightly quartz cementedsandstone in upper parts of unit A have, as indicatedby the fluid inclusion data (Figure 11), probably lostsubstantial porosity during this period.

The presence of petroleum inclusion within quartzovergrowth in the reservoir sandstones in the Ula andMime fields (Figure 11) implies that silica to someextent is mobile after oil emplacement, at least withinthe zone of immovable oil (the transition zone)(Walderhaug, 1990; Oxtoby et al., 1995). As long aspore throats and grain surfaces are water wet, disso-lution along grain contacts and stylolites may con-tinue, and the silica may diffuse through water filmson grain contacts into open pores and precipitate ascement. Furthermore, as soon as hydrocarbons enterthe pore space, the relative permeability for water isseverely reduced, and the hydrocarbon traps repre-sent dead ends for any basinwide advective porewater flow. To the contrary, isochemical cementa-tional processes within the irreducible water may stillbe active. Hence, if quartz cementation is mainly sup-ported by mass transfer by diffusion rather than byflowing pore waters, chemical compaction and quartzcementation are probably not halted by hydrocarbonemplacement.

The most convincing observation, indicating thattiming of oil emplacement into the sandstones is notthe principal cause of the porosity variations, is thefact that high-porosity sections occur in dry wells orwithin the water leg beneath the oil-water contacts(e.g., high-porosity zones in wells 2/1-6, 1/3-3, 7/11-6,and 7/11-7). The only case in which a dramatic shift inporosity is observed associated with the oil-watercontact is in well 7/11-5. However, it is possible thatthe presence of oil in the highly porous sandstones ofunit B and absence of oil in unit A is due to differencesin porosity and permeability formed prior to oilemplacement, and that oil was able to enter into theporous sandstones of unit B, but not into the tightlycemented sandstones of unit A. Hence, the porosityvariation across the oil-water contact in this well maymerely reflect a “filling down to situation” than theeffect of pore fluid on diagenesis.

From the discussion above, it appears that externalfactors such as pore pressure variations, thermal matu-rity, and oil emplacement are not the main factors con-trolling the porosity variation within the UpperJurassic sandstone reservoirs on the Cod Terrace. Oilemplacement may have retarded the porosity lossslightly, which may explain the 1%–2% higher porosi-ties within the oil zones (e.g., the difference across theoil-water contact in well 1/3-3). Original or pre-deep

Page 206: Reservoir Quality Prediction in Sand and Carbonates

burial rock characteristics are therefore more likely tocontrol the diagenetic evolution and porosity distribu-tion of the sandstones. Hence, the porosity variationmust be linked to the lithological and petrographicalcharacteristics of the different units.

Facies Control on Porosity

Two different, mutually contradictory, tendenciesof porosity variation vs. mineralogical compositionhave been identified. The observation from the welllogs that porosity appears to correlate inversely withclay content is verified by petrographical data. High-porosity intervals are in general cleaner and havelower contents of chlorite and illite and lower ClayIndex than the less-porous samples. However, it is alsoobserved that the low-porosity samples from unit C11in well 2/1-6 and from the upper part of unit A in well7/11-5 are quartz rich and contain little clay.

Three major diagenetic processes contribute to thegeneral loss of pore space with depth: mechanicalcompaction, chemical compaction, and cementation. Itis probably the varying effect of these processes on thedifferent facies that mainly causes the observed poros-ity trends. Mechanical compaction is driven by the over-burden stress (i.e., the net stress = geostatic stress –pore pressure). The bulk volume reduction is due toreorientation, cleavage, and fracturing of brittle grainsand pseudoplastic deformation of ductile grain clay

matrix. Chemical compaction is the compaction relatedto dissolution of framework grains within stylolites orwithin stressed grain contacts and is frequently associ-ated with reprecipitation of solids within adjacentopen pores. The net result is reduced pore space due tovolume reduction and cementation. Cementation isthus frequently related to chemical compaction, butmay occur following precipitation of authigenicphases following import of ions from outside. In thiscase, the porosity is decreased without any associatedbulk volume reduction, and the IGV remains constant.

Cementation

Calcite cement is present in most samples, but consti-tutes larger proportions only in thin zones. In thesezones, the intergranular volumes are large (35–45%),indicating that the precipitation occurred at relativelyshallow burial, before severe compaction. Furthermore,all extensively cemented zones are found in associationwith accumulations of bivalve fossils, whereas rem-nants of leached calcite-rich plagioclases are rare.Hence, a biogenic origin of the calcium is indicated. Inmost samples, ankerite constitutes minor proportions ofthe bulk volume (1%–5%). This cement has probablybeen formed at the expense of calcite at relatively deepburial, following release of magnesium and iron fromillitization of smectites and mixed-layer smectites. Inthe Tertiary sandstones of the Gulf Coast, Boles and

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 193

Figure 12. Pore pressure data from the Upper Jurassic reservoir sandstones from the Cod Terracearea compared to sandstones plotting along the normal porosity vs. depth trend shown in Figure 2Aand to sandstones, where high porosities appear to be related to extremely high pore pressures.

Page 207: Reservoir Quality Prediction in Sand and Carbonates

194 Ramm et al.

Franks (1979) suggested a similar origin for ankeritecement and indicated that ankerite was formed mainlyat temperatures >125°C. Note that in well 2/1-6, calciteis more abundant in the very clean sandstones of unitC11 than in the slightly more clay rich units C12 and B2,where ankerite is more abundant.

Euhedral quartz cement is the volumetrically mostimportant cement in the clean, low-porosity sandstones,but it is nearly absent in the high-porosity sand-stones. Quartz cement is normally not common inJurassic reservoirs from the Norwegian Shelf at depths<2000 m, but frequently becomes very abundantbeneath 2500–3000 m (Bjørlykke et al., 1989, 1992; Ehren-berg, 1990; Ramm and Ryseth, 1996). Fluid inclusionhomogenization temperatures also indicate that thequartz precipitation is normally initiated at burialdepths beneath 2500–3000 m (Walderhaug, 1994a).Hence, during burial toward ~2500–3000 m, the “high-porosity outliers” and the “normal-porosity sandstones”may have followed similar porosity–depth trends. Atabout this burial depth, the “normal-porosity sand-stones” started to lose more porosity by quartz cementa-tion, whereas the “high-porosity sandstones” were notcemented and remained highly porous.

Mechanical Compaction

In samples in which quartz and carbonate cementconstitute minor proportions of the rock volume,mechanical compaction seems to be the dominantmechanism of porosity destruction. Mechanical com-paction of sandstones will normally cause an exponen-tial porosity reduction with depth (Wood, 1989;Ramm, 1992), and the porosity–depth trend may beexpressed with equations on the form

(3)

where Z is depth and φis a rate factor depending on theframework strength of the rock. Equation 3 has the sameform as the nonlinear regression lines used to fitporosity–depth trends shown in Figure 6. In the result-ing porosity–depth profiles the initial, zero-depth poros-ity equals the pre-exponential factor, φ0, and αdescribesthe rate of porosity loss with depth. The nonlinearregression lines plotted in Figure 6 indicate that equa-tion 3 can be used to model the porosity evolution of thetwo end-member groups, the facies 1 (high-porosity)sandstones and the facies 3 (muddy) sandstones, byassuming φ0 = 45% and α= 0.20 and 0.49 km–1, respec-tively. The XRD data from samples that are not exten-sively quartz or carbonate cemented indicate asystematic trend of decreasing porosity with increasingclay-to-framework grain ratio (Figure 8). A regressionline, φ= 26.9 × e–3.46 ×CI, was found to describe this rela-tionship. Combining the regression line of porosity vs.Clay Index at 4100–4350 m (approximated to 4.2 km)with equation 3 and assuming φo = 45% yields:

(4)

and

(5)

Hence, a framework stability factor of 0.12 km–1 isexpected to fit the data with zero clay content. This givesa better porosity vs. depth trend than indicated by thefacies 1 (high-porosity) sandstones shown in Figure 6.Except for one sample, the maximum porosity amongthe analyzed samples from the three deeper wells is<25%. Furthermore, in Figure 6, it is observed thatporosities >25% are rare in all sandstones buried beneath4 km. Theoretical considerations on mechanical com-paction of sandstones containing varying amounts ofductile and stable framework grains suggest that whenthe content of ductile grains is low, bridging of nonduc-tile grains may prevent deformation of the ductile grains(Rittenhouse, 1971). Thus, with an ideal distribution ofsoft spherical grains, one in 21 grains could be presentwithout notable reduction in framework stability. If theclay matrix in the studied sandstones behaves similarly,it may be suggested that variations in the Clay Indexbetween 0 and 0.05 do not affect the rate of mechanicalcompaction. Accordingly, a framework stability factor of0.16 (0.12 + 0.05 ×0.82) should fit the porosity data of theclean sandstones as long as they are unaffected by chem-ical compaction and cementation. Furthermore, theporosity–depth trend of the “high-porosity sandstones”and the “poor-porosity sandstones” shown in Figure 6corresponds to the expected trend for rocks having aClay Index equal to 0.10 (0.12 + 0.82 ×0.10 = 0.20) and 0.45(0.12 + 0.82 ×0.45 = 0.49), respectively.

Chemical Compaction

Theoretical modeling of pressure solution (Ramm,1992), fluid inclusion data from varying reservoir sand-stones (Walderhaug, 1994a), and the distribution ofquartz cement in the Garn Formation off mid-Norway(Ehrenberg, 1990) and in the Brent Group and StatfjordFormation in the Northern Viking Graben (Bjørlykkeet al., 1992; Ramm and Ryseth, 1996) suggest thatporosity loss by chemical compaction and quartzcementation is important only at burial depthsbeneath 2500–3000 m. The quartz precipitation tem-peratures deduced from the fluid inclusion data fromthe Ula and Mime fields (Nedkvitne et al., 1993;Walderhaug, 1994b) indicate that quartz cementationand chemical compaction are important only beneath2500–3000 m in the Cod Terrace area as well. Dissolu-tion along grain contacts and within stylolites is theobvious source of the quartz cement, but the ampli-tude of individual stylolites (1–3 mm) and the spacingbetween stylolites in the intensively quartz cementedzones (10 cm) indicate that solution along stylolites isinsufficient in accounting for the observed amount ofcement. The IGVs in the quartz-cemented samples is21%–27%, and the sum of intergranular porosity andquartz cement is 15%–20% (Figure 9), which is lowerthan the densest packing of spherical grains. Further-more, the IGVs are lower in the quartz-cementedsandstones than the IGV in the most porous samplesfrom units C12 and B2. These observations indicatethat intergranular pressure solution has contributed,together with dissolution along stylolites, as a princi-pal source of quartz cement. This is supported by the

α = + ×

= + ×ln

4526 9

3 46 4 2 0 12 0 82.

. / . . .CI CI

φ α= × = ×× ×26 9 453 46 4 2. – . – .e eCI

φ φ α= × ×( )0 e Z–

Page 208: Reservoir Quality Prediction in Sand and Carbonates

observation that straight, concave-convex, and sutu-rated contacts between detrital quartz grains are morecommon than tangential grain contacts.

In the Ula field samples, the amounts of quartz over-growths vary 2%–8% in clean sandstone samples, butthey are rare or absent in muddy samples (Nedkvitneet al., 1993). Similar relations have previously beenobserved by Tada and Siever (1989), who reported effi-cient inhibition of quartz precipitation in sandstonescontaining more than 4% clay matrix and by Rammand Ryseth (1996), who reported particularly smallamounts of quartz cement (<3%) in samples with morethan 8% detrital clay. The samples from this unit have alower Clay Index than the adjacent porous samplesfrom units C12 and B2. Similar relationships have beenreported in the literature: Tada and Siever (1989) docu-mented efficient inhibition of quartz precipitation insandstones containing >4% clay matrix. There is, how-ever, no clear relationship between clay content andquartz cement in the clean samples from wells 2/1-6and 7/11-5; several of the quartz-cemented samplesfrom unit C11 in well 2/1-6 and from unit A in well7/11-5 have higher clay contents than some of theporous samples from adjacent sandstones.

MICROCRYSTALLINE QUARTZCOATING

Pervasive microcrystalline quartz coatings on frame-work grains have been observed in samples from wells7/11-5, 7/11-6, 7/11-10S (unit B), and 2/1-6 (units C12and B2). The samples containing this coating alwayscontain little euhedral macroquartz, and it appears thatthe coating has prevented normal quartz cementation.Similar “tiny, double-ended quartz crystals” have beenobserved in a study of Upper Jurassic sandstones in theClaymore oil field, in the British sector of the North Sea(Spark and Trewin, 1986). Those researchers noted that“The early deposition of small quartz crystals on grainsurfaces was an important factor in the porosity preser-vation. These crystals provided cement when primaryporosity was high, but only occupied a small proportionof the porosity. With further burial the crystals inhibitedthe deposition of larger pore-filling quartz and feldsparovergrowths.” If this interpretation is correct, similarrelationships can also be inferred for the Gyda andMime reservoirs, implying that the good reservoir qual-ity is a direct consequence of the microquartz coating. Touse this observation in a predictive manner, fundamen-tal questions related to its occurrence and to its ability topreserve porosity must be answered.

Occurrence

Calcedonic quartz cement is observed in samplesfrom unit B in wells 7/11-5 and 7/11-6. Precipitation ofmicro- and cryptocrystalline quartz requires initial ele-vated silica activities due to presence of amorphous sil-ica and frequently follows the generalized diageneticevolution: Opal-A—Opal-CT—cryptocrystalline—microcrystalline quartz (Williams and Crerar, 1985;Williams et al., 1985; Hendry and Trewin, 1995). Themicrocrystalline quartz coating on framework grains inthe studied well probably formed following a similar

evolution path; it is likely that the large number of smallcrystals formed by self-nucleation from a pore waterthat was highly supersaturated with respect to quartz.This self-nucleation probably occurred at a time whenan earlier opaline phase dissolved and kinetic reactionbarriers were overcome due to increased temperatures.

Similar features (microcrystalline quartz and cal-cedonic cement) have also been observed at approxi-mately 1700 m burial depth in sandstones within thevolcanic dominated Balder Formation (Eocene) inwells from block 25/11 and 25/8 in the Viking Graben(Ramm et al., 1992, unpublished Norsk HydroReports). Furthermore, recrystallized volcanic glassshards (≤1.5 cm) have been found in samples from unitB in well 7/11-5; altered sponge spicules are rathercommon in the sections containing grain-rimmingmicroquartz. Hence, the microcrystalline coatings aremost probably a consequence of initial deposition ofvolcanic or biogenic amorphous silica.

Porosity Preservation

The small microcrystalline quartz crystals are lessstable (~3%; Jahren, in press) than the normal macro-quartz, because of a larger surface energy. Hence, thatthese small crystals are preserved implies that the porewater has been supersaturated with respect to quartzever since the formation of the microquartz, probablysince burial to <1700 m (by analogy with the BalderFormation).

Precipitation of euhedral quartz overgrowths nor-mally occurs, at low degrees of supersaturation(5%–20%), by screw dislocation growth (or spiralgrowth). If physical hindrances like coatings or poison-ing atoms fixed on the growth sites are present, ahigher degree of supersaturation is required beforegrowth can commence. The crystallographic axes of thesmall (0.5–5 mm) quartz crystals are randomly ori-ented, and only occasionally will the microcrystal beoriented parallel to the underlying detrital grain.Hence, during spiral growth, precipitation at outcrop-ping dislocations will commence against the micro-quartz crystals. The alternatively two-dimensionalnucleation on flat surfaces probably does not occurbecause of the required higher degree (~35%) of quartzsupersaturation, which is not achieved during normalclay-induced quartz dissolution or pressure solutionbetween grains and in stylolites (Jahren, in press).

QUANTITATIVE POROSITY VS. DEPTH MODEL

Figure 13 illustrates a generalized quantitativemodel to explain the porosity variations in the UpperJurassic sandstones in the Cod Terrace area. Duringevaluation of new prospects, the model is applicablewhen combined with depositional models predictingthe distribution of sandstone facies (e.g., shale con-tent) and potential occurrence of zones with micro-quartz coatings.

During shallow to intermediate burial, most poros-ity reduction is due to mechanical compaction. Theporosity–depth gradient at a particular depth is

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 195

Page 209: Reservoir Quality Prediction in Sand and Carbonates

196 Ramm et al.

Figure 13. (A) Porosity depth model. (B) Model compared to the 75% He-porosity data. According to thegeneralized model, the illustrated empirical porosity–depth trends reproduce the porosity variations inthe data set. The porosity in the sandstones that are not prone to chemical compaction is expressed as φ = 45 e(–0.16 ×Z) when CI < 0.05, and φ= 45 e(–(0.12 + 0.82 ×CI) x Z) when CI > 0.05, while the porosity in the sand-stones prone to chemical compaction is expressed as φ= 45 e(–0.16 ×Z) when Z < 2.8 km, and φ= 29 – 13 ×(Z – 2.8), when Z > 2.8 km.

Page 210: Reservoir Quality Prediction in Sand and Carbonates

proportional to the existing porosity and a rate factor,depending on the framework stability. The frame-work stability is mainly a function of the clay-to-framework grain (quartz plus feldspar) ratio.Sandstones with low clay contents have high frame-work stability, and at 2500–3000 m, they still haveporosities >25%. Clay-rich sandstones and mud-stones, however, have low framework stability and, at3000 m, porosities <10% may be encountered.

At 2500 m, chemical compaction and quartz cemen-tation are initiated in clean sandstones without coat-ings on framework grains. Below ~2800 m, the rate ofporosity loss by chemical compaction becomes fasterthan the rate of porosity loss by mechanical com-paction. Empirically, these sandstones are indicated tolose ~13% porosity per kilometer burial during burialfrom ~2800 to ~4300 m. Thus, occurrences of goodreservoir quality in sandstones buried beneath 4000 mrequire high framework stability and retarded chemi-cal compaction. Inhibition of quartz cementation dueto microcrystalline quartz coatings on frameworkgrains seems to have been efficient in retarding thechemical compaction in deeply buried reservoir sand-stones in the Gyda field and in block 7/11.

CONCLUSIONS

1. The porosity variations within the Upper Jurassicsandstones from the Cod Terrace area reflect ini-tial and pre-deep burial rock composition andtexture, and are little affected by pore pressureand time/temperature variations or by timing ofhydrocarbon emplacement.

2. At shallow to intermediate burial, porosity reduc-tion is mostly by mechanical compaction. Theporosity–depth gradient at a particular depth is afunction of present porosity and framework stabil-ity. The factor found to be most influential on theframework stability of the sandstone is the clay-to-framework grain (quartz plus feldspar) ratio.

3. Below ~2800 m, two groups of relatively cleansandstones diverge rapidly with respect to poros-ity vs. depth trends. Those sandstones prone tochemical compaction and quartz cementation areseverely affected and lose ~13% porosity per kilo-meter burial between 2800–4300 m. Other sand-stones are less affected by chemical compactionand experience a slower rate of porosity loss. At4200 to 4500 km burial depth, these sandstonesstill have porosities between 20% and 25%.

4. Retarded chemical compaction in relatively cleanarenites is mainly due to inhibited quartz precip-itation and dissolution because of microcrys-talline quartz on framework grains.

5. Prior-to-drilling prediction of porosity in thesedeeply buried sandstones will be successful onlyif the inhibiting effect of the microquartz coatingon quartz cementation is recognized. Further-more, the geological evaluation should empha-size the prediction of the distribution of differentsandstone facies and the presence of sandstoneswithin the intervals deposited during periodswith high amorphous silica deposition.

ACKNOWLEDGMENTS

This paper is mainly based on research and devel-opment projects funded by Norsk Hydro, but earlyparts were conducted when M. Ramm was employedat the University of Oslo. M. Ramm would like tothank Knut Bjørlykke for criticism and suggestionsduring this period. Norsk Hydro is acknowledged forpermission to publish this study. The authors wouldalso like to acknowledge constructive reviews byCharles Curtis and Pete Turner.

REFERENCES CITED

Bjørlykke, K., P. Aagaard, H. Dypvik, D.S. Hastings,and A.S. Harper, 1986, Diagenesis and reservoirproperties of Jurassic sandstones from the Hal-tenbanken area, offshore mid-Norway, in A.M.Spencer et al., eds., Habitat of hydrocarbons on theNorwegian continental shelf: Norwegian PetroleumSociety, London, Graham & Trotman, p. 275–386.

Bjørlykke, K., M. Ramm, and G.C. Saigal, 1989, Sand-stone diagenesis and porosity modification duringbasin evolution: Geologische Rundschau, v. 78, p. 243–268.

Bjørlykke, K., T. Nedkvitne, M. Ramm, and G.C. Sai-gal, 1992, Diagenetic processes in the Brent Group(Middle Jurassic) reservoirs of the North Sea—anoverview, in A.C. Morton, R.S. Haszeldine, M.R.Giles, and S. Brown, eds., Geology of the BrentGroup: Geological Society of London Special Publi-cation 61, p. 265–289.

Boles, J.R., and S.G. Franks, 1979, Clay diagenesis inWilcox Sandstones of Southwest Texas: implicationsof smectite diagenesis on sandstone cementation:Journal of Sedimentary Petrology, v. 49, p. 55–70.

Ehrenberg, S.N., 1989, Compaction and porosity evolu-tion of Pliocene sandstones, Ventura Basin, Califor-nia: discussion: AAPG Bulletin, v. 73, p. 1274–1276.

Ehrenberg, S.N., 1990, Relationship between diagene-sis and reservoir quality in sandstones of the GarnFormation, Haltenbanken, mid-Norwegian conti-nental shelf: AAPG Bulletin, v. 74, p. 1538–1558.

Ehrenberg, S.N., 1993, Preservation of anomalous highporosity in deeply buried sandstones by grain-coating chlorite: examples from the NorwegianShelf: AAPG Bulletin, v. 77, p. 1260–1286.

Forsberg, A.W., M.B. Gowers, and E. Holtar, 1994,Multidiscipline stratigraphic analysis of the UpperJurassic strata of the Norwegian Central Trough, inA.M. Spencer, ed., Generation, accumulation andproduction of Europe’s hydrocarbons III: Proceed-ings of the European Association of PetroleumGeologists Special Publication, New York,Springer-Verlag, v. 3, p. 45–58.

Gluyas, J.G., A.J. Leonard, and N.H. Oxtoby, 1990, Dia-genesis and petroleum emplacement: the race forspace—Ula Trend, North Sea (abs.): Utrecht, Inter-national Association of Sedimentologists, 13th Inter-national Sedimentologist Congress Abstracts, p. 193.

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 197

Page 211: Reservoir Quality Prediction in Sand and Carbonates

198 Ramm et al.

Gluyas, J.G., A.G. Robinson, D. Emry, S.M. Grant, andN.H. Oxtoby, 1993, The link between petroleumemplacement and sandstone cementation, in J.R.Parker, ed., Petroleum geology of Northwest Europe:Geological Society of London, p. 1395–1402.

Hancock, N.J., and A.M. Taylor, 1978, Clay mineraldiagenesis and oil migration in the Middle JurassicBrent Sand Formation: Journal of the GeologicalSociety of London, v. 135, p. 69–72.

Hendry, J.P., and N.H. Trewin, 1995, Authigenic quartzmicrofabrics in Cretaceous turbidites: evidences forsilica transformation processes in sandstones: Journalof Sedimentary Research, v. 65, p. 380–392.

Home, P.C., 1987, Ula, in A.M. Spencer et al., eds.,Geology of Norwegian oil and gas fields: London,Graham and Trotman, p. 143–151.

Houseknecht, D.W., 1987, Assessing the relativeimportance of compactional processes and cemen-tation to reduction of porosity in sandstones: AAPGBulletin, v. 71, p. 633–642.

Jahren, J.S., 1993, Microcrystalline quartz coatings insandstones: a scanning electron microscopy studyin Karlson, ed., Extended abstracts of the 45thannual meeting of the Scandinavian Society forElectron Microscopy (abs.): SCANDEM 93.

Jahren, J.S., and M. Ramm, in press, The porosity pre-serving effects of microcrystalline quartz coatingsin arenitic sandstones; examples from the Norwe-gian Continental Shelf, in R.H. Worden and S.Morad, eds., Quartz cementation in oil field sand-stones: International Association of Sedimentolo-gists Special Publication.

Nedkvitne, T., D.A. Karlsen, K. Bjørlykke, and S.R.Larter, 1993, Relationship between reservoir diage-netic evolution and petroleum emplacement in theUla field, North Sea: Marine and Petroleum Geology,v. 10, p. 255–270.

Osborn, M., and S. Hazeldine, 1993, Evidence for reset-ting of fluid inclusion temperatures from quartzcements in oilfields: Marine and Petroleum Geology,v. 10, p. 271–278.

Oxtoby, N.H., A.W. Mitchell, and J.G. Gluyas, 1995, Thefilling and emptying of the Ula Oilfield: fluid inclu-sion constraints, in J.M. Cubitt and W.A. England,eds., The geochemistry of reservoirs: Geological Soci-ety of London Special Publication 86, p. 141–157.

Ramm, M., 1991, On quantitative mineral analysis ofsandstone using XRD: Department of Geology,Oslo, Internal Papers, v. 63, 23 p.

Ramm, M., 1992, Porosity–depth trends in reservoirsandstones: theoretical models related to Jurassicsandstones, offshore Norway: Marine andPetroleum Geology, v. 9, p. 553–567.

Ramm, M., and K. Bjørlykke, 1994, Porosity/depthtrends in reservoir sandstones: assessing the quanti-tative effects of varying pore-pressure, temperaturehistory and mineralogy, Norwegian Shelf data:Clay Minerals, v. 29, p. 475–490.

Ramm, M., and A.E. Ryseth, 1996, Reservoir qualityand burial diagenesis in the Statfjord Formation,North Sea: Petroleum Geosciences, v. 2, p. 313–324.

Rittenhouse, G., 1971, Mechanical compaction of sandscontaining different percentages of ductile grains: a

theoretical approach: AAPG Bulletin, v. 55, p. 92–96.Selley, R.C., 1978, Porosity gradients in the North Sea

oil-bearing sandstones: Journal of the GeologicalSociety of London, v. 135, p. 119–132.

Sommer, F., 1978, Diagenesis of Jurassic sandstones inthe Viking Graben: Journal of the Geological Societyof London, v. 135, p. 63–67.

Spark, I.S.C., and N.H. Trewin, 1986, Facies relateddiagenesis in the main Claymore oilfield sand-stones: Clay Minerals, v. 21, p. 479–496.

Stewart, I.J., and K. Scherverud, 1993, Structural controlson the Late Jurassic age shelf system, Ula Trend, Nor-wegian Sea, in J.R. Parker, ed., Petroleum geology ofNorthwest Europe: Geological Society of London, p. 469–483.

Tada, R., and R. Siever, 1989, Pressure solution duringdiagenesis: Annual Review of the Earth PlanetaryScience, v. 17, p. 89–118.

Walderhaug, O., 1990, A fluid inclusion study ofquartz cemented sandstones from offshore mid-Norway—possible evidence for continued quartzcementation during oil emplacement: Journal ofSedimentary Petrology, v. 60, p. 203–210.

Walderhaug, O., 1994a, Precipitation rates of quartzcement in sandstones determined by fluid inclu-sion microthermometry and temperature-historymodelling: Journal of Sedimentary Research, v. 64,p. 324–333.

Walderhaug, O., 1994b, Temperatures of quartz cemen-tation in Jurassic sandstones from the Norwegiancontinental shelf—evidence from fluid inclusions:Journal of Sedimentary Research, v. 64, p. 311–324.

Williams, L.A., and D.A. Crerar, 1985, Silica diagenesis,II. General mechanisms: Journal of SedimentaryPetrology, v. 55, p. 312–321.

Williams, L.A., G.A. Parks, and D.A. Crerar, 1985, Sil-ica diagenesis, I. Solubility controls: Journal of Sedi-mentary Petrology, v. 55, p. 301–311.

Wood, J.R., 1989, Modelling the effect of compactionand precipitation/dissolution on porosity, in I.E.Hutcheon, ed., Short Course in Burial Diagenesis:Mineralogical Association of Canada, v. 15, p. 311–362.

APPENDIX A. DEFINITION OF UPPERJURASSIC LITHOSTRATIGRAPHICAL

UNITS ON THE COD TERRACE

In the Cod Terrace area, sequence A is representedby a transgressive basal sandstone (unit A1), an openmarine upward-coarsening shale to muddy sandstone(A2), and an upper, more well sorted sandstone (A3).Sequence B is represented by a lower, fine-grainedmuddy unit (B1) and an upper, more well sorted sand-stone unit (B2). In well 1/3-3, sequence C1 is repre-sented by four units: C11 comprises clean low-porositysandstones; C12 comprises sandstones with slightlyhigher clay contents having very good reservoir qual-ity; C13 entails clean, low-porosity sandstones equiva-lent to those in unit C11; C14 comprises dark mud andfine-grained muddy sandstones. Unit C13 is present in

Page 212: Reservoir Quality Prediction in Sand and Carbonates

well 1/3-3 only, whereas unit C12 occurs in both wells1/3-3 and 2/1-6. Successively more of this sequence isabsent from wells 2/1-4, 2/1-3, and 2/1-8 due to ero-sion during the Middle Volgian unconformity. Thelithostratigraphic units C11, C12, and C13, observed inthe wells from the Gyda area, cannot be distinguishedas individual units in wells from blocks 7/11 and7/12. In those wells, unit B is overlain by a finer grainedsandstone termed unit C11–3. Unit C14, however, is recognized in all wells.

The total thickness of the Upper Jurassic sectionvaries considerably within the Cod Terrace area. Pre-Middle Volgian sections >300 m are encountered inwells 2/1-6 and 1/3-3, while sections thinner than10–20 m occur in the crestal part of the Gyda structure.This thinning across the Gyda field is mainly due toMiddle to Late Volgian erosion; only the lowermostunits A, B, and C11 are present in wells 2/1-4, 2/1-3,and 2/1-8. Initial thickness variations due to differentialsubsidence are, however, also apparent. Units A and Bare thinner in the wells from the crestal area of the Gydastructure than in the downflank wells 2/1-6 and 1/3-3.

The total thickness of the pre-Middle Volgian–UpperJurassic section is thinner in blocks 7/11 and 7/12 thanin the Gyda field area; in this area, it is practical todefine units A and B as the lithostratigraphical equiva-lents to sequences A and B. Unit A is of about uniformthickness in the wells from block 7/12, whereas unit B isthicker in the central part of the Ula field (i.e., in well7/12-2) than in the downflank wells. This relation prob-ably reflects syndepositional differential subsidence,which probably was triggered by early salt movements(Home, 1987; Stewart and Scherverud, 1993). Well7/12-5, representing a smaller structure to the north-west of the Ula field, contains a condensed unit B, indi-cating very slow subsidence during the LateKimmeridgian to Early Volgian. The condensed unitC11–3 reflects slow Early Volgian subsidence in block7/11 and 7/12 areas compared to the Gyda field area.

Along the fault edge toward the Breiflab Basin andin the northeastern part of the Cod Terrace, erosionrather than deposition was the case during the LateJurassic. Hence, Upper Jurassic sandstones are eitherabsent or much thinner than in the wells shown in Fig-ure 5 from the footwall upland (e.g., in wells 7/11-8and 7/8-4). In well 7/11-7, a 30- to 40-m-thick sand-stone is encountered between unit D and Triassicstrata. Palynological dating of this sandstone indicatesLate Middle Volgian age corresponding to sequenceC2. The sandstones probably cannot be correlated tothe other wells. An ~10-m-thick sandstone is encoun-tered in well 7/11-9 above Triassic strata; this sand-stone can probably be correlated to unit A. The UpperJurassic section is several hundred meters thick in thedownfaulted area in UK-block 23/27.

APPENDIX B. SHORT DESCRIPTION OFPETROGRAPHIC CHARACTERISTICS

OF SAMPLES FROM WELL 2/1-6

Unit C14 comprises dark, bioturbated mudstonesand fine-grained feldspathic graywackes, with occa-sional belemnite fossils. Framework grains account for50%–60%, and quartz is much more abundant thanfeldspar. Albite is more abundant than K-feldspar. Theporosity varies between 5% and 12%, mostly compris-ing microporosity within dispersed intergranularclays and matrix. Quartz overgrowths are rare, whilecarbonate cements (mainly ankerite) account for5%–10%. The matrix content (15%–30% by volume) isdominated by illite, with minor amounts of chlorite.

Unit C12 comprises fine-grained, bioturbatedarkoses. The content of framework grains is mostlyclose to 60%. Feldspars are more abundant than in theunit above and account for a fifth to a third of theframework grains. Potassium-feldspars are moreimportant than albite. The porosity varies between18% and 25%, mostly comprising intergranular macro-porosity. Quartz overgrowths are rare, while carbon-ate cements (mainly ankerite) normally account for5%–10% (volume). The clay matrix (5% by volume) isdominated by illite, with minor amounts of chlorite.

Unit C11 comprises clean, very well sorted, fine- tomedium-grained arkoses. The content of frameworkgrains is mostly close to 70% (volume). The feldsparcontent is about the same as in unit C12. The porosityvaries between 7% and 10%, mostly comprising inter-granular macroporosity. Quartz overgrowths are veryabundant, normally accounting for 10%–20%. Carbon-ate cements (calcite and ankerite) are less abundantthan in the other units (2%–3% by volume). The con-tent of clay matrix is low (2%–4% by volume) anddominated by illite.

Units B1/B2 comprise fine-grained, bioturbated,feldspathic graywackes (B1) coarsening upward intocleaner arkoses (B2). The content of framework grainsincreases from ~50% in the lower parts to ~60% (vol-ume) in the cleaner parts. The quartz and feldspardistributions are approximately the same as in unitC12. The porosity increases from <10% near the baseof the cored interval to ~20% near the top. Most of theporosity in unit B1 consists of intergranular micropo-rosity, but macroporosity dominates in unit B2.Quartz overgrowths are rare, while carbonatecements (mainly ankerite) normally account for5%–10% (volume). The clay matrix (5% to >15% byvolume) is dominated by illite, with minor amountsof chlorite.

Porosity–Depth Trends in Deeply Buried Upper Jurassic Reservoirs in Norwegian Central Graben 199

Page 213: Reservoir Quality Prediction in Sand and Carbonates

201

Chapter 13

Poroperm Prediction forReserves Growth Exploration:

Ula Trend, Norwegian North SeaJon G. Gluyas1

BP Exploración de Venezuela S.A., Edificio Centro Seguros de Sud America

Caracas, Venezuela

ABSTRACT

Much of the remaining prospectivity in the Ula trend (Norwegian CentralGraben) is deep (>3.5 km). A major risk to successful petroleum explorationin the trend is reservoir effectiveness. A few oil discoveries are not yet com-mercial because they occur in low-permeability sandstone. No simpleporosity–depth relationship exists for the whole of the Ula trend. As such,mapping of economic basement is difficult. There are, however, simpleporosity vs. depth relationships within the two main producing fields: Ulaand Gyda.

The porosity–depth relationships in the fields are due to downflankcementation by quartz. Quartz cementation was synchronous with oilemplacement, and evidence from petroleum-filled fluid inclusions has led tothe conclusion that cementing fluids and petroleum competed in a “race forspace.” The Ula trend displays evidence of all three outcomes of such a race:petroleum emplacement ahead of cementation, synchronous processes, andcementation ahead of petroleum emplacement.

Porosity prediction for undrilled prospects and prospect segments wasmade by risking the three possible outcomes of such a race for space. Thereservoir in prospect 7/12-JU4 was predicted to be oil bearing and have amean porosity of about 16.4%: a function of synchronous petroleumemplacement and cementation. The well, however, was dry. It had a meanporosity of 14%; this compares well with the predicted porosity (13.9%) atthe well location for a system in which cementation was completed beforeoil emplacement (equivalent to a porosity estimate for a dry hole).

Gluyas, J.G., 1997, Poroperm prediction for reservesgrowth exploration: Ula Trend, Norwegian NorthSea, in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 201–210.

1Present address: Monument Oil and Gas plc, London, United Kingdom.

Page 214: Reservoir Quality Prediction in Sand and Carbonates

202 Gluyas

INTRODUCTION

The Ula trend (Figure 1) of the Norwegian CentralGraben contains three producing oil fields. In decreas-ing size order they are: Ula (reserves 435 mmstb [mil-lion stock tank barrels]; Home, 1987; Brown et al.,1992), which is now in decline; Gyda (200 mmstb;Gluyas et al., 1992), which is on plateau; Mime (a fewtens of mmstb), which ceased production in 1994.Much of the exploration activity in the Ula trend todayconcentrates on reserves growth; field extensions canbe tied back to existing production facilities. Bjørnsethand Gluyas (1995) report, “remaining prospectivity inthe Ula Trend is subtle, but a large number of small tomedium sized prospects have been defined (reservesgenerally less than 100 mmstb).”

In such a mature province, the perceived explo-ration risk drives the economic viability of prospects.Even the smallest prospects can look attractive if riskis low and tieback costs are acceptable. An analysis ofdrilling statistics for the trend showed that, althoughrisks on trap presence and charging exist, they are low

relative to those associated with reservoir presenceand effectiveness (Bjørnseth and Gluyas, 1995). TheUla trend contains enough examples of lightpetroleum trapped within low-permeability rock forreservoir effectiveness to be a major concern.

This chapter examines the way an attempt wasmade to evaluate reservoir effectiveness (porosity andpermeability) of a prospect ahead of drilling.

GEOLOGICAL BACKGROUND

Well 7/12-2 was the discovery well for the Ula field(Figures 1, 2), a giant field that produces light oil froma high-quality Upper Jurassic, shallow marine sand-stone reservoir (Home, 1987; Oxtoby et al., 1995). Thetrap is a well-defined four-way-dipping dome, with>500 m of vertical closure. There are at least six oil-water contacts; the shallowest in the west is ~300 mshallower than the deepest in the east. All of the differ-ences in oil-water contacts are due to faulting ratherthan to stratigraphic effects.

ULA

ANGUS

2/4-142/2-1

7/12-5

7/8-3

MIME

GYDA

MJØLNER

ARGYLL

DUNCAN

TRYM

2/7-19

INNES

CLYDE

FULMAR

JUDY

30/1-C

ERSKINE

1/3-3

UPPER JURASSIC POOLS

"ULA TREND"

0 50km2°00'

2°00' 3°00'

3°00'

4°00'

4°00'56°00'

56°00'

57°00' 57°00'

NORWAY

DENMARK

NO

RWAY

UK

ULA - GYDA FAULT ZONE

CENTRAL TROUGH

SØRVESTLAND HIGH

7/12-2

Figure 1. Location map forthe Ula trend (fromBjørnseth and Gluyas, 1995).The 7/12-5 and 7/8-3 discov-eries remain unnamed.

Page 215: Reservoir Quality Prediction in Sand and Carbonates

Well 7/12-5 was drilled on a similar but smallerstructure in the same license a few years after 7/12-2.Depth to crest for 7/12-5 is about 400 m deeper than inthe Ula field. It too had light oil. The reservoir is thesame Upper Jurassic shallow marine sandstone presentin the Ula field. However, the average permeability ofthe reservoir in 7/12-5 (2 md) is 2 orders of magnitudelower than that for Ula (192 md). The discovery has notbeen developed or named. Deeper still is the bulk ofthe Gyda field (Block 2/1; Gluyas et al., 1992), whichlies in the adjacent license. The quality of the samereservoir in the Gyda field (40 md) is midway betweenthat of Ula and that of 7/12-5 (Figures 3, 4).

Following completion of 2-D and 3-D seismic sur-veys across the licenses in the late 1980s and early 1990s,a large number of leads and prospects were identifiedwithin the Upper Jurassic play fairway. The key risksassociated with exploration of these prospects werebelieved to be reservoir presence and permeability(Bjørnseth and Gluyas, 1995). The aim of this chapter isto illustrate an attempt to quantify one of those keyrisks—reservoir effectiveness—and to predict the reser-voir quality in prospects and prospect segments (JU2and JU4) around the 7/12-5 discovery (Figure 2). Therewas clearly no wish in the License Group to discover oilin the reservoir similar to that encountered by 7/12-5.

Poroperm Prediction for Reserves Growth Exploration: Ula Trend, Norwegian North Sea 203

Figure 2. Structural map for the7/12 block showing both the Ulafield and the prospects andprospect segments around the7/12-5 discovery.

Page 216: Reservoir Quality Prediction in Sand and Carbonates

204 Gluyas

RESERVOIR SEDIMENTOLOGY

The Ula trend sandstones are Upper Jurassic shallowmarine deposits (Figure 5). Most of the sandstones accu-mulated below fair-weather wave base as stormdeposits. Accommodation space was created by a com-bination of active rifting and associated movement ofPermian evaporites in the underlying section (Stewart,1993). The sands accumulated, base to top, as a series ofprogradational, aggradational, and finally retrograda-tional packages (Oxtoby et al., 1995) (Figure 5). Thelargely fine- to medium-grained arkosic sand was prob-ably second cycle, shed from emergent Triassic “pods”(Bjørnseth and Gluyas, 1995). As a result of the heteroge-neous development of accommodation space associatedwith salt withdrawal, reservoir thickness can vary dra-matically over short distances. For example, over a dis-tance of ~5 km in the Ula field, the reservoir thicknesschanges from 200 m to just a few tens of meters.

RESERVOIR QUALITY

Porosity, permeability, and depth data for the Ulatrend reservoir sandstone are presented in Tables 1and 2, and are plotted in Figure 3. Data are well aver-ages for the prime grainstone-texture reservoir sand-stones. Argillaceous packstone, wackestone-texture

sandstones are of very low reservoir quality every-where, and have not been included in the plots. Themost important feature of Figure 3 is that the porosityof the reservoir prime reservoir (grainstone-texture)sandstones declines dramatically from crest to flank inboth the Ula and Gyda fields. The porosity gradientsare about the same but the intercepts are different. Thefieldwide porosity–permeability relationships areplotted in Figure 4. Prospects with average permeabil-ity <10 md are probably not economically viable.

RESERVOIR DIAGENESIS

The paragenetic sequence for the reservoir sand-stones in the Ula trend is displayed in Figure 6. Thetiming of the events shown in Figure 6 is based on acombination of microthermometric determinations onfluid inclusions, radiometric age dating of illitic clay,and stable isotope analysis of carbonate cementslinked to burial history and thermal history calcula-tions. Supporting evidence for both the relative andabsolute timing of diagenetic events has been pub-lished by Gluyas et al. (1990), Gluyas and Coleman(1992), Gluyas et al. (1993), Nedkvitne et al. (1993),Oxtoby et al. (1995), and Ramm et al. (this volume).

Most of the precompaction processes have little effecton the quality of the sandstones, although local but per-vasive calcite cementation has reduced the net-gross ratein part of Gyda. The main processes to have affected thequality of the extant reservoirs are compaction andquartz cementation (Gluyas et al., 1990; Nedkvitne et al.,1993; Oxtoby et al., 1995; Ramm et al., this volume).

Of compaction and quartz cementation, quartzcementation is the key variable within individual oilfields. Figure 7 is a crossplot of quartz cement content

Figure 3. Porosity–depth plot for wells in Ula, Gyda,and 7/12-5. Data are well averages for clean grain-stones (most of the reservoir was cored in the wellson this figure). Three anomalous points are high-lighted. The relatively high porosity in 3A on thewestern flank has not been satisfactorily explained.The A-13 well on Ula lies at the north end of thefield where the sandstones shale out. The low poros-ity may be due to increased compaction associatedwith the shaly sandstone in A-13. The A-01 well onGyda has high porosity; the origin has been attrib-uted to the retardation of cementation in A-01 by oilon its migration from flank to crest of Gyda. mss =meters subsea level.

Figure 4. Mean arithmetic porosity and permeabilitydata for the Ula, Gyda, and 7/12-5 wells.

Page 217: Reservoir Quality Prediction in Sand and Carbonates

against porosity for the Ula field wells. Highly porous(and permeable) sandstones have little cement, whiletight sandstones have much cement. The relationshipbetween well-cemented sandstones of poor reservoirquality and little cemented sandstones of much betterreservoir quality is manifest in two ways. For Ulatrend fields, reservoir quality diminishes dramaticallyas a function of burial depth as quartz cement contentincreases (Figure 8). The rate of porosity decline iscommonly double the regional gradient. However,high-porosity anomalies exist. These are either thecrestal parts of oil accumulations or short intervals ofmedium-grained sandstones on the field flanks.

Two quite different explanations for the origin ofthe anomalously porous sandstones have been pub-lished. These explanations have also sought to explainthe steep porosity gradients.

Hypothesis 1: Cementation Retarded by Petroleum Emplacement

The steep cementation gradients and high-porosityanomalies within fields are postulated to be a productof the way in which cementation and petroleumemplacement interacted during filling of the trap with

oil (Gluyas et al., 1990, 1993). The hypothesis states thatthe presence of petroleum retarded cementation byquartz. The highly porous field crests were filled withoil before much cementation. The steep cementationgradients record simultaneous cementation and oilemplacement, creating a “race for space,” during whichdownward filling of the trap with petroleum progres-sively slowed cementation (Oxtoby et al., 1995).

Supporting evidence for this hypothesis is pro-vided by the abundance and distribution ofpetroleum-filled and aqueous inclusions trappedwithin the quartz cements. The crestal parts of Ulacontain only a few percentiles of quartz cement.However, this cement is quite literally full ofpetroleum-filled fluid inclusions (Oxtoby et al., 1995).The abundance of such inclusions declines down-flank and in the water leg of the field, and petroleum-filled fluid inclusions are rare. The reservoir in7/12-5 contains only aqueous inclusions. This istaken to imply that cementation was complete in thearea around 7/12-5 before oil emplacement.

Further supporting evidence for the hypothesiscomes from anomalies within the distribution ofpetroleum-filled inclusions and the distribution ofhigh-quality reservoir rock. On the eastern flank of

Poroperm Prediction for Reserves Growth Exploration: Ula Trend, Norwegian North Sea 205

Figure 5. Graphic log of theUla reservoir sandstone(from Oxtoby et al., 1995).

Page 218: Reservoir Quality Prediction in Sand and Carbonates

the field, well 7/12-7 encountered the reservoir ~400m below the field crest. The average reservoir prop-erties of the reservoir in this well fall on the field-wide porosity and permeability trends. However,the well contains ~2 m of medium-grained sand-stone at the top of the principal reservoir interval. Itis both highly porous (27%–28%) and permeable (>1darcy), having similar reservoir quality to the fieldcrest. This sand has very little quartz cement, butwhat it does contain has abundant petroleum-filledfluid inclusions. The inference at the time was thatthis particular 2-m interval represented one of the oilmigration routes into the Ula trap. Subsequent geo-chemical modeling of the source rock confirmed theeasterly filling direction for the field. A similar pat-tern exists in the Gyda field, where short intervals ofanomalously porous and permeable sandstones arepresent in western downflank well 2/1-A-01.

Hypothesis 2: Cementation Retarded by EarlyDiagenetic Precipitation of Microquartz

In this hypothesis, the presence of highly poroussandstones at depth is attributed to retardation ofquartz cementation in sandstones that have grainscoated by microcrystalline quartz. This hypothesis isfully described by Ramm et al. (this volume). Theoccurrence of microcrystalline quartz cement hasbeen linked to the distribution of relic remains ofsiliceous sponge spicules and/or volcanic glass. Thespiculitic sands accumulated as shoals, and as suchthe presence of high porosity at depth is seen to be afunction of the lithofacies distribution at the time ofsand deposition.

The two hypotheses are clearly different, and theiruse in methodologies for prediction of porosity wouldgive very different results. However, at the time thatthe Ula trend prospects were under evaluation, theexploration team working the issue accepted thehypothesis that cementation could be retarded by oilemplacement. This chapter reports their work.

Deposition

Feldspar precipitation

Calcite precipitation

Feldspar dissolution

Silica dissolution

Calcite dissolution

Compaction

Quartz precipitation

Ferroan dolomite ppt.

Illite precipitation

Stylolitization

Oil emplacement

Time (Ma)150 100 50 0

majorevents

minorevents

?

Figure 6. Summarized diagenetic history for the Ulasandstone in the Ula field. The Gyda field reservoirhas a similar history, although in Gyda calcite islocally an important cement, forming impermeablelayers. The figure is modified from Bjørnseth andGluyas (1995). Quantitative diagenetic data for theUla and Gyda wells have been published in Gluyaset al. (1990), Gluyas and Coleman (1992), Gluyas etal. (1993), Nedkvitne et al. (1993), Oxtoby et al.(1995), and Ramm et al. (this volume).

Table 1. Average Porosities for Ula Trend Wells.

Well Middepth (mss) Porosity (%)

2/1-3 –3800 22.72/1-4 –4100 19.62/1-6 –4270 9.82/1-A-01 –4100 17.92/1-A-02 –3700 21.02/1-A-04 –3840 19.62/1-A-07 –3820 21.17/8-3 –3740 13.77/12-5 –3850 11.37/12-2 –3420 19.87/12-3A –3570 21.57/12-4 –3460 18.27/12-6 –3440 20.17/12-7 –3800 14.07/12-8 –3740 15.27/12-9 –3740 16.47/12-A-01 –3670 16.47/12-A-03 –3570 17.07/12-A-03A –3685 17.07/12-A-08 –3840 10.67/12-A-12A –3509 19.77/12-A-13 –3672 10.07/12-A-13A –3540 17.07/12-A-15 –3400 19.07/12-A-18 –3480 19.2

Table 2. Field Average Porosity and Permeability.

Field Porosity (%) Permeability (md)

Ula 16.3 192Gyda 15.5 407/12-5 11.3 27/8-3 13.7 42

206 Gluyas

Page 219: Reservoir Quality Prediction in Sand and Carbonates

RESERVOIR PRESSURE

Each of Ula, Gyda, and 7/12-5 wells was overpres-sured at discovery by ~12–14 MPa at the reservoirmidpoint. Overpressure in the Ula trend reservoirs isprobably a function of burial disequilibrium caused byrapid burial in the Neogene. The whole of the area ofthe Ula trend is covered by ~2 km of Neogene mud-rich sediments. For lack of better data, the 12–14 MPaoverpressure estimate was used in the porosity calcu-lation for the prospects.

ESTIMATION OF POROSITYAND PERMEABILITY

Prediction of porosity is commonly treated as simplyan estimation of uncertainty; that is, the spread arounda most likely value. The porosity of the sandstone couldbe controlled by one of three largely distinct processes:(1) compaction alone, (2) compaction with cementation,or (3) compaction with a degree of cementationinversely controlled by petroleum emplacement. Thepossibilities were captured by risking three models. Theoutcomes for each of these models have varyingdegrees of uncertainty.

The three porosity evolution models (Figure 9) are:

• Model 1—Quartz cementation was completebefore petroleum emplacement (regional porositydecline).

• Model 2—Quartz cementation and petroleumemplacement occurred at the same time (Ulatrend porosity decline).

• Model 3—Petroleum emplacement occurredbefore quartz cementation. The sandstonesremain largely uncemented (no cementation).

Model 1—Quartz Cementation Complete Before Oil Emplacement

Many North Sea Jurassic sandstones have similarporosity–depth gradients (8% ±1% km–1) (Selley, 1978;Gluyas, 1985). Emery et al. (1993) have shown thatmany of these sandstones have only water-bearingfluid inclusions in quartz cement. They concludedthat for these sandstones, cementation by quartz wascompleted in the absence of petroleum. The sand-stones in well 7/12-5 contain only aqueous inclusionsin quartz cement.

For model 1, the prospect porosities are calculatedusing 7/12-5 data and a regional porosity gradient of–8% km–1. There are too few data to develop a waterleg gradient specifically for the Ula trend.

Model 2—Quartz Cementation Synchronous with Oil Emplacement

The Ula trend sandstones exhibit very steep poros-ity declines with depth that are associated withequally rapid increases in quartz cement with depth.The porosity–depth gradients in the Ula and Gydafields are similar, but the intercepts differ. The distri-bution of petroleum-filled fluid inclusions in thequartz cement mimics that of the porosity decline. Thehighly porous field crests with little cement contain anabundance of petroleum-filled fluid inclusions in thatcement. The well-cemented flanks of the fields showfew or no petroleum-filled fluid inclusions in theirquartz cement (Oxtoby et al., 1995).

Porosities are calculated using the 7/12-5 data forthe intercept and the Gyda/Ula data for the slope. The

Poroperm Prediction for Reserves Growth Exploration: Ula Trend, Norwegian North Sea 207

Figure 7. Relationship between quartz cement con-tent and porosity for the reservoir sandstones of theUla field. Data for both porosity and quartz cementcontent are average values for each well. The numberof samples for each porosity point is approximately100; for each quartz cement point, about 10.

Figure 8. Quartz cement vs. depth [meters subsea(mss) level] for Ula field. The two main anomaliesare at –3550 m, well 7/12-3A (Figure 3), and –3690 m,well 7/12-A12. The A12 anomaly may be due topoint-count error since it lies on the porosity–depthplot for the field.

Page 220: Reservoir Quality Prediction in Sand and Carbonates

208 Gluyas

7/12-5 data are considered a reasonable interceptbecause the prospects are clustered around this dis-covery. The application of a steep Ula trend gradientthrough the data of 7/12-5 is analogous to the Ula orGyda fields in which quartz–cemented sandstoneswithout petroleum-filled inclusions occur just abovethe deepest oil-water contacts (Oxtoby et al., 1995).

Model 3—Oil Emplacement Before Quartz Cementation

The crestal parts of both Ula and Gyda containsandstones with very little or no quartz cement. Suchsandstones are simply compacted according to theeffective burial stress (lithostatic load minus fluidoverpressure).

An estimate of porosity prior to drilling is calcu-lated using the compaction equation:

(1)

(Robinson and Gluyas, 1992), in which φ = porosity(%) and z = depth (m), with appropriate adjustmentfor an overpressure correction. An overpressure of14 MPa is equal to an effective burial depth ~1120 mless than the real burial depth using the equation ofGluyas and Cade (this volume):

(2)

in which z’ = effective burial depth (in meters), u =overpressure (in megapascals), ρr and ρw are density ofrock and water (kgm–3), g = acceleration due to gravity(ms–2), and φΣ = bulk fractional porosity of overlyingsand and mud sediment column. The porosity–depthrelationships associated with each of these models areillustrated by Figure 9.

RISKED POROSITY MODELS

The following risks were assigned on the basis ofempirical observations.

• About 1 in 20 of the Ula and Gyda wells have sig-nificant portions of their reservoir interval free ofquartz cement.

• Two fields—Ula and Gyda—had synchronous oilemplacement and cementation.

• Only 7/12-5 was cemented before oil emplacement.

There was, therefore, a nominal risk of 2:1 in favor ofsynchronous cementation and oil emplacement. How-ever, given that a significant portion of the 7/12-JU4

z zugr w

' –– –

= ( ) ( )

ρ ρ 1 φΣ

φ =+ ×

50

102 4 5 10

3

4exp–

.

–z

z

Figure 9. Modeled porosity–depth relationships. Model 1: Oil emplacement before cementation uses a regionalporosity gradient of 8% porosity loss for each additional kilometer of burial. The line passes through theporosity–depth point for 7/12-5. Model 2: Oil emplacement and simultaneous quartz cementation uses a porositygradient of 16% km–1 derived from the local Ula trend data. It too passes through the porosity–depth point forwell 7/12-5. Model 3: Oil emplacement without (before) cementation is based on the compaction curve of Gluyasand Cade (this volume), with an overpressure correction of 14 MPa. The curve is conditioned to the high-porosity,uncemented sandstones at the crest of the Ula field and downflank to similarly high-porosity sandstones in 7/12-7.

Page 221: Reservoir Quality Prediction in Sand and Carbonates

prospect was updip of 7/12-5, the chance of simultane-ous cementation and oil emplacement was estimated tobe higher. The consequent estimated risks were: model1 = 0.2, model 2 = 0.75, and model 3 = 0.05.

The porosity calculations were based on the trapconfiguration in Figure 2 and data in Table 3. Theresultant risked porosities were: JU4 = 16.4%, JU2segment C = 14.8%, JU2 segment A = 11.5%, andJU2 segment B = 14.5%.

PERMEABILITY CALCULATION

Permeabilities were calculated from the empiricalrelationship between porosity and permeability usingfield average data (Ula and Gyda, 7/12-5) (Table 4,Figure 4).

(3)

where k = permeability (in millidarcys) and φ = porosity(in percent).

The resultant risked permeabilities were: JU4 = 120md, JU2 segment C = 50 md, JU2 segment A = 4 md,and JU2 segment B = 20 md.

UNCERTAINTY CALCULATIONS

The uncertainty surrounding the porosity and per-meability data predictions was calculated using thespread of Ula field data in Figures 3 and 4. The porosityrange for the Ula field at a given depth is 6% at 95%confidence limits. Hence, 2σ on the porosity quotedabove is ±3%; this porosity variation corresponds to apermeability variation of ~0.75 magnitude (Figure 4).

WELL RESULTS

Well 7/12-10 was drilled on prospect 7/12-JU4.The reservoir was present, but the oil was missingand the well was dry. The fault system lying to thewest of the prospect is now believed to be sealing,having stopped access of petroleum to the prospect.

In consequence, the appropriate model should havebeen model 3, cementation complete before oilemplacement. The porosity predicted by this modelwas 13.9%, and that from core analysis in 7/12-10was 14.0%; this was within the confidence limits, andtherefore a perfect prediction.

DISCUSSION AND CONCLUSIONS

This approach to reservoir quality prediction mayseem sophisticated. Moreover, because this approachuses, as support, a hypothesis that is disputed, it istempting to conclude that the approach is not worth-while. However, the hypothesis is used only to explainthe porosity-to-depth relationships and not to generate amethodology for porosity prediction. The model curvesin Figure 9 are based wholly on empirical observation.

The “no cementation” curve is founded on theobservation that some of the sandstones in the Ulatrend are not cemented. The shape of the curve isbased upon experimental and empirical data (Gluyasand Cade, this volume).

The “regional porosity decline” curve of Figure 9 isbased on empirical data from the Central and NorthernNorth Sea. The sandstones in this data set are like thoseof 7/12-5 because their quartz cements do not containpetroleum-filled fluid inclusions (Emery et al., 1993).

The “Ula trend porosity decline” is the local curve.The porosity gradient is steep when compared withregional data; the component sandstones that make upthe Ula trend porosity decline are distinct insofar asthey contain petroleum trapped in inclusions in quartzcement. Thus, the three models for porosity predictioncan be used without reference to a hypothesis toexplain the models.

In conclusion, the lack of a simple porosity-to-depthrelationship for the Ula trend as a whole drove investi-gations to reveal how porosity was destroyed. This inturn delivered a methodology that allowed better useof the empirical porosity–depth data for reservoirquality prediction.

log . – .10 0 36 3 80k = φ

Poroperm Prediction for Reserves Growth Exploration: Ula Trend, Norwegian North Sea 209

Table 3. Parameters for Ula Trend Prospects.

Depth JU2 JU2 JU2 (m) JU4 Segment C Segment A Segment B

Depth to crest 3440 3440 3720 3540Depth to closure 3525 3900 3846 3846Mid-volume depth 3510 3675 3820 3675

Table 4. Predicted Porosities for Ula Trend Prospects.

Porosity JU2 JU2 JU2(%) JU4 Segment C Segment A Segment B

Model 1 13.9 13.1 Proven 11.5 12.7Model 2 16.5 14.6 — 14.0Model 3 25.7 25.2 — 25.0

Page 222: Reservoir Quality Prediction in Sand and Carbonates

210 Gluyas

ACKNOWLEDGMENTS

I wish to thank Per Svela, Grete Block-Valge, andPer Christian Mjelde for helping improve this manu-script. I also thank BP Norge and partners for givingme permission to publish this work.

REFERENCES CITED

Bjørnseth, H.M., and J.G. Gluyas, 1995, Petroleumexploration in the Ula trend, in S. Hanselein, ed.,Petroleum exploration in Norway: Norsk Petrole-umsforening/NPF, Special Publication 4, Proceed-ings of the Norwegian Petroleum Conference,December 9–11, 1991, Stavanger, Norway, Elsevier,Amsterdam, p. 85–96.

Brown, A., A.W. Mitchell, I.R. Nilssen, I.J. Stewart, andP.T. Svela, 1992, Ula field: relationship betweenstructure and hydrocarbon distribution, in B.T.Larsen and R.M. Larsen, eds., Structural and tec-tonic modelling and its application to petroleumgeology: Norsk Petroleumsforening/NPF, SpecialPublication 1, Elsevier, Amsterdam.

Emery, D., P.C. Smalley, and N.H. Oxtoby, 1993, Syn-chronous oil migration and cementation in sand-stone reservoirs demonstrated by quantitativedescription of diagenesis: Philosophical Transactionsof the Royal Society of London, v. 344, p. 115–125.

Gluyas, J.G., 1985, Reduction and prediction of sand-stone reservoir potential, Jurassic North Sea: Philo-sophical Transactions of the Royal Society ofLondon, v. A315, p. 187–202.

Gluyas, J.G., K. Byskov, and N. Rothwell, 1992, A year inthe life of Gyda production: IBC, Advances in Reser-voir Technology, London—Conference Proceedings,p. 187–202.

Gluyas, J., and C.A. Cade, this volume, Prediction ofporosity in compacted sands, in J.A. Kupecz, J.Gluyas, and S. Bloch, eds., Reservoir quality predic-tion in sandstones and carbonates: AAPG Memoir69, p. 19–28.

Gluyas, J.G., and M.L. Coleman, 1992, Material flux andporosity changes during diagenesis: Nature, v. 356,p. 52–53.

Gluyas, J.G., A.J. Leonard, and N.H. Oxtoby, 1990,Diagenesis and petroleum emplacement: the race

for space—Ula Trend, North Sea (extended abs.):Nottingham, England, 13th International Sedimen-tological Congress, International Association ofSedimentologists, Utrecht, p. 193.

Gluyas, J.G., A.G. Robinson, D. Emery, S.M. Grant, andN.H. Oxtoby, 1993, The link between petroleumemplacement and sandstone cementation, in J.R.Parker, ed., Petroleum geology of NorthwestEurope: Barbican, London, Geological Society ofLondon Proceedings of the 4th Conference, p. 1395–1402.

Home, P.C., 1987, The Ula oilfield block 7/12, Norway,in A.M. Spencer et al., eds., Geology of theNorwegian oil and gas fields: Norwegian PetroleumSociety, London, Graham & Trotman, p. 143–152.

Nedkvitne, T., D.A. Karlsen, and K. Bjørlykke, 1993,Relationship between diagenetic evolution andpetroleum emplacement: Marine and PetroleumGeology, v. 10, p. 225–270.

Oxtoby, N.H., A.W. Mitchell, and J.G. Gluyas, 1995,The filling and emptying of the Ula oilfield (Norwe-gian North Sea), in J.M. Cubitt and W.A. England,eds., The geochemistry of reservoirs: GeologicalSociety Special Publication 86, p. 141–158.

Ramm, M., A.W. Forsberg, and J.S. Jahren, this volume,Porosity-depth trends in deeply buried Upper Juras-sic reservoirs in the Norwegian Central Graben: anexample of porosity preservation beneath the nor-mal economic basement by grain-coating micro-quartz, in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones and car-bonates: AAPG Memoir 69, p. 177–199.

Robinson, A.G., and J.G. Gluyas, 1992, Model calcula-tions of sandstone porosity loss due to compactionand quartz cementation: Marine and PetroleumGeology, v. 9, p. 319–323.

Selley, R.C., 1978, Porosity gradients in North Sea oil-bearing sandstones: Journal of the Geological Soci-ety of London, v. 135, p. 119–132.

Stewart, I.J., 1993, Structural controls on the Late Juras-sic age shelf system, Ula trend, Norwegian NorthSea, in J.R. Parker, ed., Petroleum geology of North-west Europe: Barbican, London, Geological Societyof London Proceedings of the 4th Conference, p. 469–484.

Page 223: Reservoir Quality Prediction in Sand and Carbonates

211

Chapter 14

Predicting Porosity Distribution WithinOolitic Tidal Bars

Larry J. CavalloStonewall Gas Company

Jane Lew, West Virginia, U.S.A.

Richard SmosnaDepartment of Geology and Geography, West Virginia University

Morgantown, West Virginia, U.S.A.

ABSTRACT

The Mississippian Greenbrier Limestone is a major gas reservoir in theAppalachian basin, but its complex porosity patterns often deter active explo-ration. In southern West Virginia, the reservoir consists of oolitic tidal barsthat are composites of smaller shoals. Porosity trends closely follow the ooid-grainstone facies that occupied shoal crests where coarse-grained, well-sortedooid sand was generated with either unidirectional or bidirectional cross-beds. Nonporous packstone occurred in adjacent tidal channels, and a transi-tional grainstone/packstone facies of marginal porosity was situated alongthe flanks of the shoals. The key to drilling successful wells is in understand-ing the complex internal geometry of Greenbrier ooid shoals. A well penetrat-ing the oolite with good porosity and bimodal cross-beds should be offsetperpendicular to the dip directions; that is, parallel to the shoal axis.However, a well penetrating thin, porous limestone with one dominant cross-bed azimuth should be offset opposite to that dip direction; that is, up theflank of the ooid shoal. Shaly interbeds characterize the edges of the shoalsand mark the limit of productive wells. Schlumberger ’s FormationMicroScanner log, which provides data on both lithology and cross-bedding,has proven to be a useful tool in predicting the distribution of oolite porosity.

INTRODUCTION

Oolitic reservoirs in the Mississippian GreenbrierLimestone (Union Member; Figure 1) have histori-cally produced significant quantities of natural gasacross the central Appalachian basin. Exceptionalwells in West Virginia, for example, will ultimately

produce 1–2 billion ft3 (3–6 ×107 m3) and occasionally≤9 billion ft3 (27 ×107 m3) of gas. Nevertheless, a com-plex geological setting with seemingly randomporosity patterns, rapid facies changes, and involveddiagenetic histories has deterred active exploration.Gas production varies widely, depending on paythickness (15–97 ft; 4.6–30 m), porosity (3%–28%),

Cavallo, L.J., and R. Smosna, 1997, Predicting porositydistribution within oolitic tidal bars, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 211–229.

Page 224: Reservoir Quality Prediction in Sand and Carbonates

212 Cavallo and Smosna

and permeability (0.01–15 md). Well spacing in pro-ducing fields, although not restricted, averages 2400 ft(730 m), and typical treatment of the pay zone consistssimply of acidization with 15% HCl.

Only recently have productive zones within theUnion Member been fit into a regional depositionalmodel (Kelleher and Smosna, 1993). Union oolites ofsouthern West Virginia were deposited as a belt oftidal bars positioned along a northeast–trending hingeline in the Greenbrier gulf (Figure 2). Along this hingeline (or series of hinge lines), a shallow-marine shelf tothe northwest dropped off into a somewhat deeper

basin to the southeast (Donaldson, 1974; Kelleher andSmosna, 1993). Strong tidal currents striking the shelfedge are thought to have generated north-west–trendingoolitic bars and intervening channels, similar to thosethat border Tongue of the Ocean and Exuma Sound,Bahamas (Ball, 1967; Halley et al., 1983). Kelleher andSmosna (1993) delineated by isopach maps eight tidalbars in McDowell, Wyoming, Raleigh, and Mercer coun-ties (Figure 3); four central bars were well defined byprevious drilling, but the existence of the outer bars wasat that time somewhat speculative. More recent drillingillustrates that the belt of tidal bars does in fact continuealong trend to the northeast and southwest.

This study concentrates primarily on the drillingresults of Stonewall Gas Company in the northeastern-most oolitic bar, named Blue Jay from the lease nameof its discovery well (Figure 3). Ten successful wells(out of 13) have been drilled on the southern terminusof that tidal bar. The geometry and makeup of the BlueJay bar, however, is more complicated than originallybelieved. Our purpose in this chapter is to refine Kelle-her and Smosna’s (1993) model so that it accuratelyreflects the highly variable nature of porosity develop-ment within the reservoir. In this way, porosity trendscan be better predicted, leading to new discoveriesand an effective exploitation of these oolitic bars. Totest our refined model and illustrate its usefulness, weapply concepts developed for the Blue Jay bar toanother bar where drilling is ongoing. At present,eight wells have been drilled on the southern terminusof the Poca Land bar, situated 10 km to the west (Fig-ure 3), and an additional five wells are scheduled forthe near future. Locations for new wells in the PocaLand bar will be based on geological predictions fromthe more detailed depositional model.

Our analyses make full use of Schlumberger’s Forma-tion MicroScanner (FMS) log, a relatively new technol-ogy that accurately measures minute differences in rockresistivity (Serra, 1989). Computer processing of FMSresistivities produces a color image of the inside of thewell bore (resembling a core photograph); darker huesrepresent more conductive elements or beds, and lighterhues more resistive elements or beds. More than a sim-ple dipmeter, the FMS tool allows continuous observa-tion of detailed lateral and vertical changes in rockproperties. From the shapes and patterns revealed onthe FMS image, an experienced geologist can interpretthe rock texture (in this study, oolites, shale partings,mottling, and interbedding of lithologies), sedimentarystructures (cross-beds, bed thickness, the nature of bed-ding contacts, and stylolites), and structural features(regional dip, fractures). Lithologic identifications basedon FMS logs are confirmed by petrographic analysis of12 sidewall cores recovered from the Union Member.

GEOLOGICAL SETTING

The Mississippian Greenbrier Limestone accumu-lated during a major transgression of an epeiric sea intothe Appalachian foreland basin. Deposition occurred ina broad gulf that extended across parts of sixAppalachian states (Figure 2). The gulf was bordered by

Figure 1. Stratigraphic members of the GreenbrierLimestone and lowest Bluefield Formation pickedon the gamma-ray log of well 118, Mercer County,West Virginia.

Page 225: Reservoir Quality Prediction in Sand and Carbonates

the Acadian highlands to the east, exposed lowlands tothe north, and the Cincinnati Arch to the west (deWittand McGrew, 1979). Northern lowlands shed small vol-umes of terrigenous sediment into the basin as paralicsandstones that interfinger with sandy marine lime-stones, while minor red shales represent an eastern allu-vial plain that rimmed the Acadian highlands (Adams,1970; Brezinski, 1989; Carney and Smosna, 1989;Smosna and Koehler, 1993). Paleomagnetic data placethe basin ~10° south of the Equator (Scotese, 1984). Theclimate may have been fairly arid due to a rain-shadoweffect behind the Acadian highlands, leading to reducedrunoff and a low level of terrigenous input from sur-rounding landmasses (Cecil, 1990).

During rapid basin subsidence, cherty skeletalwackestones, packstones, and mudstones accumulatedin relatively deep-water environments of the south-eastern basin (basal Hillsdale and Denmar members;Figure 1). Red silty shales (overlying Taggard Member)mark a brief progradation of the eastern alluvial plain.Subsidence then slowed, the sea transgressed to itsgreatest limits, and ooid grainstones formed on theshallow shelf (Pickaway and Union members). Theselatter oolites are the focus of our study. Finally, anincrease in terrigenous sediments (Alderson Member)signifies the close of carbonate deposition. The overly-ing Mauch Chunk Group (including the Lillydale Shaleand Reynolds Limestone members of the Bluefield For-mation) consists of marine and fluvial-deltaic shalesand sandstones with only minor limestones.

A number of Mississippian structural/stratigraphichinge lines have been identified across West Virginiaand Kentucky on the basis of regional and local isopachmapping (Flowers, 1956; Donaldson, 1974; MacQuownand Pear, 1983; Carney and Smosna, 1989; Kelleher andSmosna, 1993). In each case, the formation thickensmarkedly over a short distance. In the area of this study,for instance, the Greenbrier Limestone thickens to thesoutheast at a rate of 6.5 m/km north of the hinge lineand 8.9 m/km south of the line. The formation attainsits maximum thickness of ~900 m in neighboring Vir-ginia, and it is thought that basin subsidence may havetaken place along deeply seated normal faults beneaththese down-to-the-south hinge lines (Donaldson, 1974;MacQuown and Pear, 1983). Rapid deposition was ableto keep pace with the differential subsidence, so eventhe basin center remained relatively shallow.

COMPOSITE BARS

A stratigraphic cross section ~3 km long has beenconstructed along the axis of the Blue Jay bar usinggamma-ray and bulk-density well logs (Figure 4). Thestratigraphic interval extends from the ReynoldsLimestone Member, a marker bed across the entirestate, down to a thin dolomite near the middle of theUnion Member that serves as a local datum. Thin andshaly in the south (well 80), the Union oolite becomesthicker and less argillaceous to the north (well 145).Furthermore, three distinct units within the oolite can

Predicting Porosity Distribution Within Oolitic Tidal Bars 213

Figure 2. Paleogeography ofthe Greenbrier gulf in thecentral Appalachian basin.A belt of tidal bars in theUnion oolite member waspositioned along a hingeline that separated the basinon the south from the broadshelf to the north.

Page 226: Reservoir Quality Prediction in Sand and Carbonates

214 Cavallo and Smosna

be identified on gamma-ray well logs in the south,although they merge northward into a single blockysignature. Interbeds of shale also occur on the easternand western margins of the Blue Jay bar. Presence orabsence of shale interbeds thus provides some indica-tion of a well’s position on the tidal bar.

The isopach map of the Union oolite, ranging inthickness from 9 m on the bar’s eastern edge to 19 malong its axis, delineates a north-south orientation(Figure 5). The bar is shown to terminate south of theexisting wells for two reasons. First, the gamma-raylog for well 80 (Figure 4) shows the oolite to be rathershaly and to have lost its blocky signature. It is pre-dicted, therefore, that its thickness decreases markedlytoward the south. Second, this southern termination ofthe bar closely coincides with the southern termina-tion of other bars to the west (Figure 3) (Kelleher andSmosna, 1993). A line connecting these southern termi-nations parallels the trend of the tidal-bar belt and thepostulated hinge line; presumably, water depth wastoo great and tidal currents too weak for substantialooid formation basinward of this line.

Quite apparent on the cross section of Figure 4 is thevariable nature of porosity development along the bar,

where porosities ≥6% (density <2.60 g/cm3) are noted onthe bulk-density logs. Porosity is present in the lowestunit of wells 80 and 151 only, in the lower two units ofwell 159, in the upper two units of well 133, in the upperunit of well 145 only, and in all three units of well 118.

An isopach map of the lower unit is depicted in Fig-ure 6A. Shaly deflections in the gamma-ray well logsdetermine the base and top of this unit, although thetop becomes more difficult to distinguish in the centralarea of the bar. Varying in thickness from 2 to 8 m, theunit appears as two laterally linked shoals. The shoalsare ~1500 m wide and 3200 m long; they are of equalsize and shape, both reveal a parallel northwest-south-east orientation, and a thin intershoal area separatesthe two. This pattern may continue along the entirelength of the Blue Jay bar, perhaps 30 km to the north.

An isopach map of the middle unit indicates a thick-ness range between 3 and 9 m (Figure 6B). Two shoalsare similar to those of the lower unit in terms of size,shape, orientation, and intershoal thin. Shoals of themiddle unit are situated immediately above the inter-shoal areas of the lower unit, and shoals of the lowerunit are capped by the intershoal area of the middle unit.The isopach map of the upper unit (Figure 6C) illustratesa single curvilinear shoal. This shoal is 1200 m wide,3600 m long, and 2–7 m thick. A smaller shoal of proba-ble limited extent and based on only one well may belocated to the east. The curvilinear shoal lies directlyabove the combined crests of shoals identified in thelower and middle units.

Figure 7 shows porosity-isopach maps constructedfor the lower, middle, and upper units. These mapsdepict stratigraphic thicknesses with a porosity ≥6%.Trends in porosity-thickness correspond closely to thetotal thickness of each unit (Figures 6, 7). Porosity inUnion grainstones consists of intercrystalline micro-pores between calcitic microrhombs that make up theindividual ooids; primary interparticle porosity hasbeen occluded by calcite cement. Kelleher and Smosna(1993) suggested that growth of the bars above sealevel or periodic lowstands enabled meteoric water toenter and diagenetically alter the ooids to calciticmicrorhombs, thereby creating secondary microporos-ity within the grains. Comparable interpretations forthe recrystallization of porous ooids in other limestonereservoirs have been offered by Keith and Pittman(1983) and Ahr (1989). Close agreement between strati-graphic thickness of the Union oolitic units and theirporosity thickness supports this hypothesis: crests ofthe shoals where sedimentation was greatest wouldhave stood higher and been exposed longer to theinfiltration of meteoric water and would now possessthe greatest porosity.

LITHOFACIES

The FMS image logs provide lithologic informationfor the Blue Jay bar. On the basis of rock texture (inparticular, oolites, shale partings, mottling, andinterbedding of lithologies) and sedimentary struc-tures (cross-beds, thickness of bedding or bed sets, and

Figure 3. Outline of north-west–trending tidal barsand intervening channels in the Union Member, asmarked by the 9-m isopach contours (from Kelleherand Smosna, 1993). The two detailed study areas ofthis chapter include the southern ends of the BlueJay bar in Mercer County and the Poca Land bar inWyoming and Raleigh counties.

Page 227: Reservoir Quality Prediction in Sand and Carbonates

Predicting Porosity Distribution Within Oolitic Tidal Bars 215

Figu

re 4

. Cro

ss s

ecti

on A

A’ a

lon

g th

e ax

is o

f th

e B

lue

Jay

tid

al b

ar. G

amm

a-ra

y (G

R) a

nd

bu

lk-d

ensi

ty (B

D) w

ell l

ogs

show

thre

e d

isti

nct

por

ous

un

its

wit

hin

the

ooli

te o

f th

e U

nio

n M

emb

er, a

nd

str

atig

rap

hic

inte

rval

s w

ith

>6%

por

osit

y ar

e m

ark

ed in

bla

ck o

n th

e b

ulk

-d

ensi

ty lo

gs. T

he

low

est u

nit

(dar

k s

tip

pli

ng)

is p

rese

nt i

n w

ells

159

, 118

, 151

, an

d 8

0; th

e m

idd

le u

nit

(med

ium

sti

pp

lin

g) in

wel

ls 1

59,

133,

an

d 1

18; a

nd

the

up

per

un

it (l

igh

t sti

pp

lin

g) in

wel

ls 1

45, 1

33, a

nd

118

. Lin

e of

sec

tion

is m

ark

ed o

n F

igu

re 5

.

Page 228: Reservoir Quality Prediction in Sand and Carbonates

216 Cavallo and Smosna

reactivation surfaces), three lithofacies have beendefined: cross-bedded ooid grainstone, bioturbatedpackstone, and a transitional grainstone/packstone.

Cross-Bedded Ooid Grainstone

Evident on FMS images for the lower unit of well151 (Figure 8) is a grainy, pockmarked texture, a char-acteristic feature of this lithofacies. Although resolu-tion of the FMS tool is not fine enough to observeindividual ooids, the range of resistivities presentamong the ooids accounts for the diagnostic graininess.Equally characteristic of the facies is an abundance ofwell-defined cross-bed sets. Images reveal these sets asa series of stacked sinusoidal curves with the same dipazimuth and magnitude. Dip angles range from 2° to29°, averaging 16° (regional structural dip is <2°). Fre-quently associated with the cross-beds are reactivationsurfaces, relatively flat-lying surfaces that separatecross-bed sets (Klein, 1977; Smosna and Koehler, 1993).Cross-beds above and below these erosional surfacesoften display vastly different dip azimuths and magni-tudes. These bed sets, reaching a maximum thicknessof 60 to 70 cm, represent sand waves or megaripplesthat migrated across the surface of the ooid shoals.

The amount of shale partings is low in the ooid-grain-stone facies. Shale, less resistive than surrounding lime-stone, appears darker on the images. Where present,shales are thinner than 5 cm. They are usually at the topor bottom of the unit and outside the porous pay zones.

Six sidewall cores recovered from the Union oolitein well 145 (Figure 5) are of this lithofacies. Thin-section analysis establishes these rocks as ooid grain-stones. Ooids account for >80% of the total grains andhave a mean grain size of 0.65 mm (coarse sand). Mostooids possess a thick cortex around a nucleus of fos-sils, peloids, or small intraclasts, although several ofthe largest ooids have a relatively thin coating. Thesesix sidewall cores have a mean porosity of 9.3% andmean permeability of 0.145 md.

Bioturbated Packstone

This lithofacies displays a mottled texture on theFMS logs (Figure 9), a texture vastly different fromthat of the ooid grainstone. Unfortunately, no sampleshave been retrieved for thin-section analysis. A pack-stone lithology, however, is considered most likely,because in modern carbonate settings, channelsbetween oolitic tidal bars are floored by burrowed,muddy pelletal sand (Ball, 1967; Harris, 1979). In ananalogous manner, mottled patches with varyingresistivity resulted from churning and mixing of theGreenbrier sediment by burrowing infauna.

Cross-bedding and other sedimentary structuresare not observed in FMS images. Instead, drilling-induced fractures are a diagnostic feature of this litho-facies, an indication of the packstone’s lower physicalstrength. Where not extensively bioturbated, beddingis on the order of a few centimeters. Stylolites are acommon feature, appearing on the image as thin,highly irregular traces of less-resistive minerals. Also,the number of interbedded shales is relatively high.Rocks of this facies have no appreciable porosity andare not of reservoir quality.

Transitional Grainstone/Packstone

The third lithofacies is a combination of the previoustwo, containing characteristics of both the cross-bed-ded ooid grainstone and the bioturbated packstone(Figure 10), but subtle differences justify designatingthese rocks as a separate facies. Its grainy texture is lessextensive than that of the ooid grainstone, due to sig-nificant interbeds of mottled packstone and shale.Cross-beds are common but not pervasive, and seldomare they truncated by reactivation surfaces. Drilling-induced fractures are absent in the burrowed-mottledsections of this facies. Six sidewall cores from well 145show the rock type to be an ooid grainstone. In contrastto ooids of the pure grainstone lithofacies, these have afiner grain size (0.45 mm, or medium sand) and rela-tively thin coatings. Moreover, ooids account for only50%–80% of the total grains; uncoated fossils, peloids,and intraclasts are abundant.

Porosity in the transitional facies is typically <4%and of marginal reservoir quality. Cored samples havea mean porosity of 2.7% and permeability of <0.01 md.

Lithofacies Distribution

Figure 11 depicts the geographical extent of thethree Blue Jay lithofacies during deposition of the

Figure 5. Isopach map of the Union oolite delineatesa north-south orientation for the southern termina-tion of the Blue Jay tidal bar. Contour intervalequals 3 m.

Page 229: Reservoir Quality Prediction in Sand and Carbonates

Predicting Porosity Distribution Within Oolitic Tidal Bars 217

Figure 6. Isopach maps of the lower unit (A), middle unit (B), and upper unit (C) illustratethe individual ooid shoals that make up the Blue Jay tidal bar. Arrows indicate main dipazimuths of cross beds (regional structural dip is negligible). Contour interval equals 1.5 m.

Page 230: Reservoir Quality Prediction in Sand and Carbonates

218 Cavallo and Smosna

Figure 7. Porosity-isopach maps (stratigraphic thickness with ≥6% porosity) of the lower unit(A), middle unit (B), and upper unit (C) correspond closely to the total thickness of each unit.Compare with Figure 6. Contour interval equals 3 m.

Page 231: Reservoir Quality Prediction in Sand and Carbonates

Predicting Porosity Distribution Within Oolitic Tidal Bars 219

Figure 8. An FMS image log of an ooid grainstone,demonstrating its grainy texture as well as cross-bedsets (stacked sinusoidal curves with the same dipazimuth and magnitude). Reactivation surfaces atdepths 3139.4, 3140.5, and 3142.0 ft abruptly terminatethe cross bedding. Lower unit of Union oolite, well151, Mercer County.

Figure 9. An FMS image log of a bioturbated pack-stone, demonstrating its mottled texture as well asshale interbeds (thick dark horizontal bands), stylo-lites (thin dark horizontal bands at depths 3115.5,3118.1, 3120.2, and 3120.6 ft), and vertical fractures.Upper unit of Union oolite, well 151, MercerCounty.

Page 232: Reservoir Quality Prediction in Sand and Carbonates

220 Cavallo and Smosna

Figure 10. An FMS image log of the transitional grainstone/packstone, demonstrating a mot-tled texture [depths 3240–3243 ft and 3247–3249 ft] and a grainy texture [depths 3243–3247 ftand 3249–3254 ft]. Middle unit of Union oolite, well 143, Mercer County.

Page 233: Reservoir Quality Prediction in Sand and Carbonates

Predicting Porosity Distribution Within Oolitic Tidal Bars 221

Figure 11. Facies maps of lower unit (A), middle unit (B), and upper unit (C) of the Unionoolite at Blue Jay bar. Each shoal consists of a central cross-bedded oolite-grainstone faciessurrounded by a transitional grainstone/packstone facies. Burrowed packstone occupies theadjacent tidal channels. Circled wells are those with FMS logs.

Page 234: Reservoir Quality Prediction in Sand and Carbonates

222 Cavallo and Smosna

lower, middle, and upper units. Comparisons of thesemaps to unit isopach maps (Figure 6) and porosity-isopach maps (Figure 7) illustrate a clear relationshipamong lithofacies, shoal thickness, and porosity.

Cross-bedded ooid grainstone occurs where unitthickness and porosity are greatest, corresponding tothe crest of shoals atop the Blue Jay tidal bar. The well-washed, well-sorted, coarse-grained ooids represent ahigh-energy setting in which water depth did notexceed 2 m (Newell et al., 1960; Ball, 1967; Harris, 1979).The thickness of cross-bed sets generally increasesupward within each stratigraphic unit. Bed sets arethickest (maximum ~70 cm) at the base, reflecting thegreatest water depth, and decrease to ~20 cm near thetop as sediments aggraded into shallower water (Ball,1967). At the very top of the unit, bedding again thick-ens to ~60 cm; perhaps these thicker cross-bed sets rep-resent beach ridges or dunes that capped theshallowing-upward sequence (Halley et al., 1983).

Bioturbated packstone is positioned on the lowerflanks of the shoals and in adjacent tidal channels, wherethe unit is thin and nonporous. Water depth exceeded4–5 m, and the energy level was drastically lower.

Transitional grainstone/packstone is situatedaround and between ooid shoals, and inside the bound-ary of the tidal bar. In this position, tidal-current inten-sity fluctuated between the high energy of ooidgrainstone (shoal crest) and the low energy of biotur-bated packstone (adjacent channel), and the two end-member lithologies became interbedded. The greaternumber of uncoated grains, finer grain size, thinneroolitic coating, and shale/packstone interbeds collec-tively suggest a shoal environment of slightly deeperwater and less favorable for ooid formation. On AndrosIsland and Joulters Cays, in the Bahamas, modern andPleistocene ooid shoals can similarly be divided intotwo subenvironments: a central, very oolitic area sur-rounded by an area of lower ooid concentration (Harris,1979; Boardman et al., 1993). In the Mississippian Ste.Genevieve Limestone of Indiana, Zuppann (1993)described a very oolitic lithofacies in the central portionof an ooid shoal, with a decreasing percent of ooidsdown the flanks.

PALEOCURRENT ANALYSIS

The FMS image logs also depict cross-bed dip direc-tions for the Union oolite as a whole and for each ofthe three units. These data, displayed by arrows on theseveral isopach maps, are then used to infer paleocur-rent directions at the time of deposition. Throughoutthe oolite, cross-beds have a consistent bimodal dipazimuth: northeast-southwest and separated by ~180°.This bimodal pattern confirms a tidal influence duringconstruction of the bar (Klein, 1977; Smosna andKoehler, 1993).

Cross-bed dip directions in the unit subdivisions(Figure 6) define three scenarios: (1) wells drilled on ornear the shoal axis exhibit bimodal cross-beds with180° separation of dips, (2) wells on the flanks of theooid shoals exhibit cross-beds that generally dip away

from the shoal axis, and (3) wells located outside ashoal exhibit a dip pattern that may or may not indi-cate the position of any nearby shoal.

Figure 12, a schematic cross section, portrays thethree scenarios and their relationship to shoal thickness.Recognition of these relationships becomes vital whenusing the FMS image log as a predictive tool to aid inthe placement of offset wells. A well penetrating theUnion oolite with good porosity and 180° bimodalcross-dips should be offset perpendicular to the cross-bed dip directions; that is, parallel to the shoal axis. Awell penetrating a thin, porous limestone section withone dominant cross-bed dip direction should be offsetin the direction opposite to that dip direction; that is, upthe flank of the ooid shoal. And a well penetrating lime-stone with no porosity (that is, outside the shoal) mayexhibit a dip pattern unrelated to the shoal’s position.

SEDIMENTARY MODEL

The origin of Union tidal bars can be explained fol-lowing Ball’s (1967) interpretation of modern bars inthe Bahamas. Greenbrier tidal currents may haverepeatedly swept over the hinge line separating thedeeper basin to the southeast from the shallow shelf tothe northwest. Ocean water was perhaps rapidly dis-placed with each tidal change, and the interfacebetween shelfal and basinal water masses becameunstable. Splitting into a number of equally spaceddigits, these fingerlike flows were then responsible forthe oolite-belt geometry (Ball, 1967). In shallow-marine environments with strong reversing tidal cur-rents, sand accumulates in zones of shear where, forpart of the daily cycle, adjacent currents flow in oppo-site directions (Swift and Niedoroda, 1985). Unionbars probably began their development in such a man-ner, forming in areas between the digitate currents.Ooid sand deposited in the shear zones retarded thetidal flows, resulting in additional sand deposition,which further retarded the tidal flows. Due to this con-tinuing feedback between sea-floor topography anddiminishing current strength, the bars grew upwardthrough time (Swift and Niedoroda, 1985). Theseprocesses ultimately produced a broad belt of alternat-ing bars and channels. The bars have a somewhat sin-uous axis, and the southern termination of the Blue Jaybar, in particular, trends north-south (Figures 3, 5).Tidal currents here are inferred to have flowed north-south through the intervening channels.

Flow over the Blue Jay bar itself was, however,oblique to its axis (that is, northeast-southwest), as indi-cated by the position of individual shoals (Figure 13).This oblique flow direction resulted from refraction oftidal currents where they approached the bar: the cur-rents turned through 45° as they passed over the flank ofthe bar and upward toward the crest. The ooid shoals,the most active portions of the tidal bar, were alignedcrosswise to this flow. Furthermore, the currents musthave accelerated as they moved up the shoals, due to therapidly decreasing cross-sectional area of the water col-umn (Swift, 1985). Flow along the shoal crest was

Page 235: Reservoir Quality Prediction in Sand and Carbonates

reduced by friction in shallow water, and when passingover the crest, tidal flow decelerated further, where thewater column expanded on the downcurrent flank. Theorientation of individual shoals reveals that flood cur-rents dominated the western margin of the Blue Jay bar,and ebb currents, its eastern margin.

At the beginning of a tidal cycle, ooids were erodedfrom the upcurrent (southwestern) flanks of the shoalsby accelerating, north-directed flood currents (Figure13). Erosion and entrainment of sand produced thenumerous reactivation surfaces observed on FMSimage logs. In contrast, ooids were deposited on thecrest and downcurrent (northeastern) flank from thedecelerating flood currents. Sand here accumulated inthe form of sand waves or megaripples, with cross-beds that dip northeastward. Halley et al. (1983) docu-mented the presence of sand waves, similarly oriented45° to the long axis, that adorn the crests of moderntidal bars at Tongue of the Ocean, Bahamas. With areversal of the tide, south-directed ebb currents—refracted 45° to a more southwesterly direction—eroded ooids from the (now) upcurrent northeastern

flank, producing reactivation surfaces. Grains weretransported to the crest and to the southwesterndowncurrent flank, and deposited in cross-beds dip-ping to the south-southwest.

Frequently reversing tides were therefore responsi-ble for the characteristic dip patterns of the FMS logs.Cross-beds on the crest exhibit a bidirectional dip with180° separation; on the flanks, cross-beds are unidirec-tional and dip away from the shoal’s crest. As anexample, the FMS image in Figure 8 depicts typicalsedimentary structures that formed near the axis of anooid shoal. This 7-ft (2-m) stratigraphic interval occurswithin the lower oolite unit of well 151, MercerCounty (see Figure 11A for location). A set of cross-beds may be observed at depth 3140.5–3141.4 ft, with adip direction of N35E produced by the flood current.The set has been truncated above by a reactivation sur-face (depth 3140.5 ft), in turn, is overlain by another setof cross-beds (depth 3139.4–3140.5 ft) S25W and pro-duced by the ebb current. Other sets of northeast- andsouthwest-dipping cross-beds, as well as additionalreactivation surfaces, are visible in Figure 8.

Predicting Porosity Distribution Within Oolitic Tidal Bars 223

Figure 12. Schematic crosssection of an ooid shoal,indicating internal sedi-mentary structures, thick-ness of pay zone, and FMScross-bed dip directions.Inset depicts position ofthis cross section relative toshoal’s axis.

Page 236: Reservoir Quality Prediction in Sand and Carbonates

224 Cavallo and Smosna

The Blue Jay bar is not a single, homogeneousdeposit; it is composed of three vertically stacked shoals.These shoals have an average thickness of 8 m, width of1.5 km, length of 3.5 km, and spacing of 2.1 km. Similarto composite bars described by Evans (1970), Klein(1977), and Zuppann (1993), we assume that migratingooid shoals of the Union Member merged through timeinto the larger bars. Each generation of shoals, however,is offset somewhat from those beneath (Figures 6, 11).Those of the lower unit built upward to sea level byrapid sand deposition, at which time sea-floor topogra-phy caused a lateral shift in the tidal currents. Sites ofooid deposition moved to the intershoal areas. Shoals ofthe middle unit then grew upward to sea level, forminga continuous bar composed of two generations ofsmaller shoals. With a slight relative rise of sea level,additional accommodation space was created for theshoal in the upper unit. This upper shoal lies across thecrests of those underlying it, demonstrating that its posi-tion was governed by remnant topography of the com-posite lower/middle bar. In addition, the curvilinearextension of the upper shoal toward the northeast (Fig-ure 6C) resembles the Holocene spitlike buildup thatevolved around the open northern margin of Joultersooid shoal, where ooids are swept along by longshoretransport (Harris, 1979; Boardman et al., 1993). At theclose of Union time, a major sea level rise brought ooiddeposition to an end across the region (Carney, 1993).

POROSITY PREDICTIONS

The Poca Land study area lies two tidal bars awayfrom Blue Jay. Development drilling using FMS image

logs has just begun, and a bit of artistic license or geo-logical intuition enters into the construction of the sev-eral maps. However, the refined depositional modeldeveloped for the Blue Jay bar serves as a valuableguide in predicting the distribution of porosity at PocaLand and in locating additional wells.

The Union oolite thins to the southeast, reflecting asouthward termination of the tidal bar (Figure 14). Shaleinterbeds in the oolite of the southernmost four wellssuggest that they are situated near the bar’s margin.Well 403, in contrast, centrally located along the axis,exhibits a blocky gamma-ray signature. Termination ofthe Poca Land bar near well 155 matches the southerntermination of other bars in the Greenbrier tidal-bar belt(Figure 3). As discussed above, water depth north of thisline proved ideal for the formation of ooids, but was pre-sumably too great south of the line. Thickness variationsalong the bar axis (~15 m in wells 403 and 1304, 6.7 m inwell 1384, and 7.0 m in well 155) result from differentialstacking of the constituent ooid shoals.

Gamma-ray and bulk-density well logs allow theoolite to be subdivided into three informal units (Figure15). An isopach map of the lower unit (Figure 16A)depicts two laterally linked, north-south shoals of equalshape and size (2–8 m thick, 1.6 km wide, and 2.4 km

Figure 13. Interpreted regime of dominant tidal cur-rents responsible for constructing ooid shoals of theBlue Jay bar.

Figure 14. Isopach map of the total Union oolite delin-eates a northwest-southeast orientation for the PocaLand tidal bar. Contour interval equals 3 m. Due tomechanical problems, well 186 was never completed.

Page 237: Reservoir Quality Prediction in Sand and Carbonates

Predicting Porosity Distribution Within Oolitic Tidal Bars 225

Figu

re 1

5. C

ross

sec

tion

BB

’ alo

ng

the

axis

of

the

Poc

a L

and

bar

. Gam

ma-

ray

(GR

) an

d b

ulk

-den

sity

(BD

) wel

l log

s sh

ow th

ree

dis

tin

ct u

nit

s w

ith

inth

e oo

lite

of

the

Un

ion

Mem

ber

; str

atig

rap

hic

inte

rval

s w

ith

>6%

por

osit

y ar

e m

ark

ed in

bla

ck o

n th

e B

D lo

gs. T

he

low

est u

nit

is p

orou

s (m

ediu

mst

ipp

lin

g) in

wel

ls 4

03, 1

304,

137

8, a

nd

138

4; th

e m

idd

le u

nit

is p

orou

s (l

igh

t sti

pp

lin

g) in

wel

ls 4

03 a

nd

130

4; th

e u

pp

er u

nit

has

no

por

osit

y. L

ine

ofse

ctio

n is

mar

ked

on

Fig

ure

14.

Page 238: Reservoir Quality Prediction in Sand and Carbonates

226 Cavallo and Smosna

long), separated by an intershoal thin. A third shoal isprojected to the south—in the wide gap between wells1384 and 155. As with the Blue Jay bar, these mappedshoals may persist to the northwest along the entire30-km length of the Poca Land tidal bar. The middleunit (Figure 16B) also consists of two shoals with similarorientation, size, shape, and intershoal thin. Placementof a third shoal near well 155 is somewhat awkward,because it is unusually far from the two known shoals.Perhaps two smaller shoals could be projected hereinstead of one. As observed on isopach maps for theBlue Jay bar, constituent shoals in the Poca Land barappear to be offset: shoals of the middle unit are situ-ated immediately above the intershoal areas of thelower unit, and shoals of the lower unit are capped bythe intershoal area of the middle unit. The upper unit ofthe Poca Land bar occurs as a thin blanket of 1–3 m(isopach map not included) without any ooid shoals.

Porosity-isopach maps for the lower and middleunits (Figure 17) correspond closely to the unit isopachmaps (Figure 16). The trends of greatest porosity par-allel crests of the ooid shoals. The interpretation is thesame as for the Blue Jay bar: shoal crests stood higherabove the surrounding sea floor; they were subse-quently exposed longer when relative sea level

dropped, infiltrating meteoric water created themicroporosity within ooids; consequently, areas ofgreatest porosity match stratigraphic thicks (Kelleherand Smosna, 1993). Ooid shoals of the lower unit havethe best developed porosity; limestone thickness withporosity >6% ranges ≤4.6 m. Porosity in shoals of themiddle unit occurs only in two northern wells, wherethe maximum thickness of porous limestone is 3.7 m.No porosity has been encountered in the upper unit.

Lithofacies of the Poca Land bar have been inter-preted based on five FMS image logs. A lithofacies mapof the lower unit (Figure 18A) illustrates that cross-bedded ooid grainstones occupy the crest of the centralshoal; this facies is projected to the other two shoalswhere the unit is porous and thick. The transitionalgrainstone/packstone facies surrounds the shoal crestsand extends into slightly deeper water. No wells actu-ally penetrated the burrowed packstone of the adjacenttidal channels; existence of this lithofacies is postulatedfrom the Blue Jay sedimentary model.

The cross-bedded grainstone facies has yet to befound in the middle unit (Figure 18B). However, thefact that all five logged wells fall within the transi-tional facies strongly suggests that the grainstone mustbe nearby. Furthermore, well 1304 contains 0.6 m of

Figure 16. Isopach maps of the lower unit (A) and middle unit (B) illustrate the individual ooid shoals that makeup the Poca Land tidal bar. Arrows indicate main dip azimuths of cross beds. Contour interval equals 1.5 m.

Page 239: Reservoir Quality Prediction in Sand and Carbonates

limestone with 6% porosity, hinting that the ooid grain-stone facies must lie immediately to the north. In a sim-ilar manner, the grainstone facies is projected to well403 on the northernmost shoal (not logged with FMS),based on the greater unit thickness and higher porosity.

The upper unit consists solely of burrowed pack-stone (lithofacies map not included). In contrast to theBlue Jay bar, water depth at Poca Land during deposi-tion of the upper unit was presumably too great for themaintenance of ooid shoals because (1) the Greenbriershelf may have had a slight tilt, deepening somewhatto the southwest from the Blue Jay bar (Kelleher andSmosna, 1993) with (2) a rise of sea level near the end ofUnion deposition (Carney, 1993). No ooid grainstonesor transitional grainstone/packstones are expected inthe upper unit of the Poca Land study area.

Cross-bed dip directions for the total Union ooliteof the Poca Land bar (Figure 16) reveal a curious pat-tern when compared to those of Blue Jay. Instead ofdisplaying bimodal north-south dip azimuths, dipsare almost consistently to the south (southwest tosoutheast). The northern component does not gener-ally exist. A lack of bimodal dips does not invalidatethe idea of tidal construction of the bars; the fivelogged wells are interpreted to be situated on the

southwest side of the tidal bar, where tidal currentswould have been south-directed.

Cross-bed dips in the lower unit (Figure 16A) showall three scenarios discussed for the Blue Jay bar. Well1378, situated on the axis of the central shoal, exhibitsa 180° separation (northeast-southwest) in dip direc-tions. Well 1304 shows southwest dips indicative of aposition on the western flank of the same shoal, andwell 1384 shows southern dips indicative of a positionon its southern nose. Wells 1380 and 1382, positionedoff the shoals, have dip patterns not influenced by thenearby shoal. Cross-bed dips in the middle unit (Fig-ure 16B) lie outside of the main shoal bodies, but theirconsistent southward dips give evidence of an ooidshoal north of well 1304. The FMS image logs illustraterandom dip patterns in the nonporous packstones ofthe upper unit (map not included).

Using the information for the Poca Land study area,we can predict the locations for future successful wells.Three new wells in Raleigh County (indicated by thestars in Figure 18A) are anticipated to penetrate thecentral ooid shoal of the lower stratigraphic unit. Thisprediction is based on a combination of (1) the thickillustrated on our isopach map, (2) a local maximum inporosity thickness, (3) the proximity of other wells that

Predicting Porosity Distribution Within Oolitic Tidal Bars 227

Figure 17. Porosity-isopach maps (stratigraphic thickness with >6% porosity) of the lower unit (A) and middleunit (B), Poca Land tidal bar. Contour interval equals 1.5 m.

Page 240: Reservoir Quality Prediction in Sand and Carbonates

228 Cavallo and Smosna

penetrated ooid grainstone, and (4) cross-beds in adja-cent wells indicative of a flank position. Two additionalwells in Raleigh County are anticipated to encounterthe northern ooid shoal, but this prediction is less cer-tain. We chose these locations based on the isopachthick and a local maximum in porosity thickness.Lastly, two future wells in Wyoming County may dis-cover a southern shoal in the lower Union unit. Thisprediction is made solely by extending the trend ofooid shoals into the area between wells 1384 and 155with the same spacing established to the north.

Applying the same reasoning, we predict the loca-tions for three successful wells in the middle Unionoolite (indicated by the stars in Figure 18B). One wellshould penetrate the central ooid shoal, based again on acombination of (1) the thick illustrated on our isopachmap, (2) a local maximum in porosity thickness, (3) theproximity of other wells that penetrated ooid grain-stone, and (4) cross-beds in adjacent wells indicative of aflank position. A second well will encounter the north-ern ooid shoal, a prediction supported by the isopachthick and a local maximum in porosity thickness. Athird well may discover a southern shoal, but this final

prediction is the least hopeful. The location is proposedsolely by continuing the trend of ooid shoals from thenorthwest; however, spacing between these shoalsremains uncertain.

CONCLUSIONS

The oolitic tidal-bar belt in the Union Member of theGreenbrier Limestone is not a continuously porousbody. Rather, it is composed of three stratigraphic unitsthat may contain highly porous shoal facies, marginallyporous transitional facies, and nonporous channelfacies. The three-dimensional geometry of thesenatural-gas reservoirs is quite complex: contemporane-ous shoals within a single tidal bar were laterallylinked; moreover, they stacked with a vertical offset asthey grew through time. Their orientation is oblique tothe general trend of the bar, a result of refraction of thetidal currents responsible for their development. Shoals,measuring 8 m thick, 1500 m wide, and 3500 m long,consist of ooid grainstone. On FMS logs, this lithofaciesdisplays a grainy texture, abundant cross-bed sets, andreactivation surfaces. Cross-beds along the crest have abimodal dip direction perpendicular to the shoals’ axes,

Figure 18. Facies maps of lower unit (A) and the middle unit (B) of the Union oolite at Poca Land bar. Circledwells are those with FMS logs; stars are the selected locations for future wells.

Page 241: Reservoir Quality Prediction in Sand and Carbonates

whereas those on the flanks are unimodal and directedaway from the shoals’ axes. The thickest pay zones(≤3 m of grainstone with >6% porosity) occur along theshoals’ crests. Shale partings become common aroundthe shoal margins where the ooid grainstone passesthrough a transitional lithofacies into nonporous biotur-bated packstone of the adjacent tidal channel.

ACKNOWLEDGMENTS

The authors acknowledge the reviews of JulieKupecz, Laura S. Foulk, and Neil Hurley, whose com-ments and suggestions improved the manuscript.Stonewall Gas Company provided the data and gavepermission to publish the results. Alison Hanham andDebbie Benson drafted the illustrations.

REFERENCES CITED

Adams, R.W., 1970, Loyalhanna Limestone—cross-bedding and provenance, in G.W. Fisher, F.J. Petti-john, J.C. Reed, and K.N. Weaver, eds., Studies ofAppalachian geology—central and southern: NewYork, Interscience Publishers, p. 83–100.

Ahr, W.M., 1989, Early diagenetic microporosity in theCotton Valley Limestone of East Texas: SedimentaryGeology, v. 63, p. 275–292.

Ball, M.M., 1967, Carbonate sand bodies of Florida andthe Bahamas: Journal of Sedimentary Petrology, v. 37,p. 556– 591.

Boardman, M.R., C. Carney, and P.M. Bergstrand,1993, A Quaternary analog for interpretation of Mis-sissippian oolites, in B.D. Keith and C.W. Zuppann,eds., Mississippian oolites and modern analogs:AAPG Studies in Geology 35, p. 227–241.

Brezinski, D.K., 1989, Late Mississippian depositionalpatterns in the north-central Appalachian basin,and their implications to Chesterian hierarchalstratigraphy: Southeastern Geology, v. 30, p. 1–23.

Carney, C., 1993, The drowning of ooid shoals: Missis-sippian Greenbrier Limestone near the West Vir-ginia dome, in B.D. Keith and C.W. Zuppann, eds.,Mississippian oolites and modern analogs: AAPGStudies in Geology 35, p. 141–148.

Carney, C., and R. Smosna, 1989, Carbonate depositionin a shallow marine gulf, the Mississippian Green-brier Limestone of the central Appalachian Basin:Southeastern Geology, v. 30, p. 25–48.

Cecil, C.B., 1990, Paleoclimate controls on stratigraphicrepetition of chemical and siliciclastic rocks: Geology,v. 18, p. 533–536.

deWitt, W., and L.W. McGrew, 1979, The Appalachianbasin region, in L.C. Craig and C.W. Connor, eds.,Paleotectonic investigations of the MississippianSystem in the United States: U.S. Geological SurveyProfessional Paper 1010, p. 13–48.

Donaldson, A.C., 1974, Pennsylvanian sedimentation ofthe central Appalachians, in G. Briggs, ed., Carbonif-erous of the southeastern United States: GeologicalSociety of America Special Paper 148, p. 47–78.

Evans, W.E., 1970, Imbricate linear sandstone bodiesof Viking Formation in Dodsland-Hoosier area ofsouthwestern Saskatchewan, Canada: AAPG Bul-letin, v. 54, p. 469–486.

Flowers, R.R., 1956, A subsurface study of the Green-brier Limestone in West Virginia: West Virginia Geo-logical & Economic Survey, Report of InvestigationNo. 15, 17 p.

Halley, R.B., P.M. Harris, and A.C. Hine, 1983, Bankmargin, in P.A. Scholle, D.G. Bebout, and C.H. Moore,eds., Carbonate depositional environments: AAPGMemoir 33, p. 463–506.

Harris, P.M., 1979, Facies anatomy and diagenesis of aBahamian ooid shoal: Sedimenta VII, University ofMiami, Florida, 163 p.

Keith, B.D., and E.D. Pittman, 1983, Bimodal porosity inoolitic reservoir—effect on productivity and logresponse, Rodessa Limestone (Lower Cretaceous),East Texas basin: AAPG Bulletin, v. 67, p. 1391–1399.

Kelleher, G.T., and R. Smosna, 1993, Oolitic tidal-barreservoirs in the Mississippian Greenbrier Group ofWest Virginia, in B.D. Keith and C.W. Zuppann,eds., Mississippian oolites and modern analogs:AAPG Studies in Geology 35, p. 163–173.

Klein, G.D., 1977, Clastic tidal facies: Champaign, Illi-nois, Continuing Education Publication Co., 149 p.

MacQuown, W.C., and J.L. Pear, 1983, Regional andlocal geologic factors control Big Lime stratigraphyand exploration for petroleum in eastern Kentucky:Kentucky Geological Survey, Series XI, SpecialPublication 9, p. 1–20.

Newell, N.D., E.G. Purdy, and J. Imbrie, 1960, Bahamianoolitic sand: Journal of Geology, v. 68, p. 481–497.

Scotese, C.R., 1984, Paleozoic paleomagnetism and theassembly of Pangea, in R. Van der Voo, C.R. Scotese,and N. Bonhommet, eds., Plate reconstruction fromPaleozoic paleomagnetism: American GeophysicalUnion, Geodynamic Series, v. 12, p. 1–10.

Serra, O., 1989, Formation MicroScanner image interpre-tation: Houston, Schlumberger Educational Services,117 p.

Smosna, R., and B. Koehler, 1993, Tidal origin of a Mis-sissippian oolite on the West Virginia Dome, in B.D.Keith and C.W. Zuppann, eds., Mississippian oolitesand modern analogs: AAPG Studies in Geology 35, p. 149–162.

Swift, D.J.P., 1985, Response of the shelf floor to flow,in R.W. Tillman, D.J.P. Swift, and R.G. Walker, eds.,Shelf sands and sandstones: SEPM Short CourseNotes 13, p. 135–241.

Swift, D.J.P., and A.W. Niedoroda, 1985, Fluid and sedi-ment dynamics on continental shelves, in R.W. Till-man, D.J.P. Swift, and R.G. Walker, eds., Shelf sandsand sandstones: SEPM Short Course Notes 13, p. 47–133.

Zuppann, C.W., 1993, Complex oolite reservoirs in theSte. Genevieve Limestone (Mississippian) at Fol-somville field, Warrick County, Indiana, in B.D. Keithand C.W. Zuppann, eds., Mississippian oolites andmodern analogs: AAPG Studies in Geology 35, p. 73–89.

Predicting Porosity Distribution Within Oolitic Tidal Bars 229

Page 242: Reservoir Quality Prediction in Sand and Carbonates

231

Chapter 15

Predicting Reservoir Quality at theDevelopment Scale: Methods for Quantifying

Remaining Hydrocarbon Resource inDiagenetically Complex Carbonate

ReservoirsR.P. Major

Mark H. HoltzThe University of Texas at Austin, Bureau of Economic Geology

Austin, Texas, U.S.A.

ABSTRACT

The Jordan (San Andres) reservoir comprises ~400 ft (120 m) of upward-shoaling subtidal to peritidal carbonate strata, which is now thoroughlydolomitized and partly cemented by sulfates. Subtidal facies include domi-nant pellet packstone/grainstone, with local bryozoans, algae, and coral bio-herms and associated skeletal grainstone flanking beds. The lower part ofthe subtidal section is characterized by stratigraphically distinct zones inwhich permeability has been enhanced by a postburial carbonate-leachingevent. These diagenetically altered (leached) zones crosscut subtidal deposi-tional facies. Peritidal facies are nonporous mudstone and generally non-porous pisolite packstone characterized by abundant sulfate cement. Thepisolitic rocks are locally porous and permeable where sulfate cement iseither leached or absent from fenestrae.

Cumulative production is 68 million stock tank barrels (MMSTB) of 218MMSTB original oil in place, which is a recovery efficiency of 31%. A total of47 MMSTB of remaining mobile oil occurs as bypassed oil in the contactedupper part of the reservoir, which has been penetrated by well bores; 12MMSTB of mobile oil is in the uncontacted lower part, which has not beenpenetrated by well bores. The most prospective areas for increased produc-tion by waterflood profile modification in the contacted part of the reservoirare the southwest corner of the field, where low-permeability, diageneticallyunaltered subtidal rocks are incompletely swept, and the eastern central partof the field, where heterogeneous permeability in peritidal rocks has resultedin an incomplete sweep. The most prospective areas for increased production

Major, R.P., and M.H. Holtz, 1997, Predicting reser-voir quality at the development scale: methods forquantifying remaining hydrocarbon resource indiagenetically complex carbonate reservoirs, inJ.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones andcarbonates: AAPG Memoir 69, p. 231–248.

Page 243: Reservoir Quality Prediction in Sand and Carbonates

232 Major and Holtz

INTRODUCTION

A major challenge for predicting reservoir quality atthe development scale is having a sufficiently detailedunderstanding of the geometry and extent of individualflow units (sensu Ebanks, 1987) within a reservoir. Ourknowledge of depositional facies patterns and geome-tries in sedimentary rocks is based on extensive docu-mentation of modern sediments, sedimentary processes,and ancient rocks exposed in outcrops. Our ability topredict depositional facies relationships at a scale that ismeaningful in maturely developed reservoirs is fairlyadvanced; our ability to predict diagenetic patterns thatcontrol reservoir quality is, however, at a much moreprimitive stage. We present here a case study in whichdiagenetic alteration of a carbonate reservoir controlsflow-unit geometry and, based on subsurface mapping,the geometry of these diagenetically controlled flowunits can be used to predict reservoir quality and toquantify remaining resource. The Guadalupian (UpperPermian) San Andres Formation of the Permian Basin,West Texas and southeastern New Mexico, provides anopportunity to test new reservoir characterization andresource assessment techniques.

The general depositional facies tracts of San Andresreservoirs are divided into four categories: inner ramp,ramp crest, outer ramp, and slope/basin. Flow units inouter ramp facies tract reservoirs may be controlled inlarge part by postdepositional diagenetic alteration ofrelatively homogeneous depositional facies (Ruppel etal., 1995). In this chapter, we review the geologic andengineering parameters that control reservoir quality,and the volume and distribution of remaining oil, in amature, outer ramp San Andres reservoir—the JordanSan Andres reservoir on University of Texas Lands(University Lands) in Ector and Crane counties, Texas.

GEOLOGIC SETTING ANDPRODUCTION HISTORY

The paleogeography of the Permian Basin was con-trolled by Pennsylvanian tectonism that deformedPrecambrian basement and pre-Pennsylvanian sedi-mentary rocks (Galley, 1958; Ward et al., 1986). Dur-ing the Permian, sedimentation in the region occurredin two basins, the Delaware Basin on the west and theMidland Basin on the east, separated by thesouth-southeast–trending Central Basin Platform(Figure 1). The Central Basin Platform was the site of

shallow-water ramp carbonate sedimentation,whereas the central portions of the Delaware andMidland basins were the sites of siliciclastic deposi-tion (Galley, 1958; Ward et al., 1986).

The Permian stratigraphic section on the CentralBasin Platform contains Wolfcampian, Leonardian,and Guadalupian shallow-water carbonate strata,many now thoroughly dolomitized, and includes rela-tively minor zones of siliciclastic-rich carbonates.Guadalupian carbonates are in conformable and grada-tional contact with overlying Ochoan evaporites andsiliciclastic red beds deposited during increasinglyrestricted marine conditions in the Permian Basin.

Jordan field is one of a complex of five fields—Pen-well, Jordan, Waddell, Dune, and McElroy, termed thePJWDM field complex (Major et al., 1988)—that pro-duce from both San Andres and Grayburg reservoirs(Longacre, 1980, 1983; Harris et al., 1984; Bebout et al.,1987; Major et al., 1988; Harris and Walker, 1990). Jor-dan field produces from a San Andres reservoirlocated on a low-relief, broad anticlinal structure witha northwest–trending axis (Figure 1). The structurewas created by drape of Permian sediments overburied Pennsylvanian faults that trend oblique to theapproximate eastern margin of the Central Basin Plat-form (Ward et al., 1986).

The San Andres reservoir at Jordan field is com-posed of dolomitized rocks exhibiting textures indica-tive of sediments deposited in subtidal, open-marineenvironments that shoaled upward to tidal-flat envi-ronments. These facies prograded from west to eastacross the platform, and the tidal-flat section thickenswestward. This westward thickening of low-porosityand low-permeability tidal-flat facies provides anupdip seal, and oil production is mainly from the east-ern flank of the broad anticline.

San Andres reservoirs in the Permian Basin can becategorized into four facies tracts: (1) inner ramp, (2)ramp crest, (3) outer ramp, and (4) slope/basin (Ker-ans et al., 1994; Ruppel et al., 1995). The relatively dis-tal setting of outer ramp reservoirs, such as Jordanfield, results in relatively low depositional faciesdiversity. In this setting, minor fluctuations in rela-tive sea level did not result in exposure to shoaling,higher energy environments, or subaerial exposure.Thus, although the upper part of the Jordan SanAndres reservoir represents shoaling to tidal-flatdepositional environments, much of the reservoir iscomposed of subtidal, open-marine facies that have a

through well-bore deepening into the uncontacted part of the reservoir arethe southeast corner of the field, where high-permeability, diageneticallyaltered subtidal rocks are uncontacted, and the central part of the field, wherehigh-permeability, diagenetically altered subtidal rocks are uncontacted. Anunderstanding of diagenetically controlled reservoir properties can be used topredict the locus of remaining resource and to design recovery strategies.

Page 244: Reservoir Quality Prediction in Sand and Carbonates

low depositional texture diversity. In this deposi-tional setting, the influences of postdepositional dia-genetic alteration can play a major role in control ofreservoir flow units (Ruppel et al., 1995).

Jordan field is, in terms of discovery and develop-ment history, typical of San Andres reservoirs of theCentral Basin Platform. It is a very mature field, hav-ing been discovered in 1937. Typical early wells hadinitial potential flow rates of several hundred to 1000bbl of oil per day. In the late 1940s, annual productionreached 1.9 million stock tank barrels (MMSTB).Annual production declined through the 1950s and1960s to a low of 1.0 MMSTB. A program of infilldrilling, well deepening, and conversion of producingwells to water injection wells began in 1969, followingperipheral waterflooding in 1968. By 1971, a modifiedfive-spot waterflood was in place, and annual produc-tion peaked in 1975 at 2.2 MMSTB. Annual productionsteadily declined through the late 1970s and early1980s and is now ~650 thousand stock tank barrels(MSTB). The present well spacing is ~20 acres per well,and the two Jordan field units on University Landshave a cumulative production of 68 MMSTB.

STUDY AREA AND AVAILABLE DATA

The study area is The University of Texas Landspart of Jordan field, which comprises ~4500 acres andis 66% of the field. Although some form of wireline-logdata are available for nearly all wells in the study area,the majority of logs are neutron or density-neutron

logs. As is discussed below, the most useful porositytool in this reservoir is the acoustic log, and most ofthese were run in the 1970s during infill drilling forconversion to waterflood. Virtually all resistivity logsare post-1970. Thus, all resistivity data are postwater-flood and, because flooding is assumed to have sub-stantially changed the resistivity of interstitial porewaters, these resistivity logs cannot be used to reliablycalculate fluid saturations.

Seven conventional cores, generally 300–400 ft(90–120 m) long, are available from within the studyarea (Figure 1). These were augmented by two Jordanfield cores immediately west of the University Landsboundary and 14 cores in the East Penwell San AndresUnit, which offsets Jordan field to the north (Major etal., 1990). All but two of the cores from Jordan fieldhave been analyzed for porosity and permeabilityusing high-temperature analytic techniques. As is dis-cussed in detail below, the presence of gypsum in thisreservoir requires more expensive, more time-consuming, low-temperature core analysis for accu-rate porosity and permeability measurements. Thus,although there are numerous cores for lithologicdescription, there are relatively few reliable core-derived porosity and permeability data.

LITHOLOGIC RESERVOIRDESCRIPTION

The Jordan San Andres reservoir is interpreted tohave been deposited in the outer ramp facies tract. As

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 233

Figure 1. Permian Basin paleogeography during San Andres (middle Guadalupian) time and loca-tion of Jordan field. The inset is a structure map of the University Lands part of the Jordan (SanAndres) reservoir.

Page 245: Reservoir Quality Prediction in Sand and Carbonates

234 Major and Holtz

discussed above, this location on the ramp margin ofthe Central Basin Platform results in relatively lowdepositional facies diversity (Ruppel et al., 1995), andpostdepositional diagenetic alteration can play animportant role in variations in petrophysical proper-ties. Lithologic description of the reservoir is dividedinto depositional facies and diagenetic overprint; thegoal of lithologic description is to divide the reservoirinto flow units (sensu Ebanks, 1987), which is a criticalfirst step for predicting reservoir quality.

Depositional Facies

The San Andres reservoir at Jordan field is assignedto the outer ramp facies tract of Ruppel et al. (1995)because it is dominantly composed of rocks depositedat or below fair-weather wave base. These open-marine rocks are overlain by rocks deposited in a tidal-flat setting during a period of relative sea levellowstand. Accordingly, the depositional faciesdescribed here are divided into two parts.

Open-Marine Depositional Facies

Open-marine facies are pellet packstone/grainstoneand bioherms composed of bryozoans, algae, andcorals, with associated flanking facies of skeletal grain-stone. Calcium sulfate cements are common. The pelletpackstone/grainstone facies, which is the volumetri-cally dominant reservoir facies, is composed of variableamounts of mud matrix and spherical to ovoid fecal pel-lets ~0.2–0.5 mm in diameter. Fossils of open-marineinvertebrates are common, especially fusulinids and

bivalves. Burrow structures are rare, and there is a gen-eral lack of laminations due to thorough bioturbation.Fecal pellets were deposited as soft carbonate mud andexhibit a wide range in degree of preservation, as ischaracteristic of many modern low-energy settings(Wanless et al., 1981). The pellets in this thoroughlydolomitized rock are commonly not visible on slabbedcore surfaces. Thus, these rocks may be incorrectlydescribed as mudstone or, where skeletal grains areabundant, as wackestone. Where pellets are well pre-served, the rock has interparticle porosity; where pelletshave been destroyed by compaction, porosity is lowand is generally intercrystalline, moldic, or both. Exten-sive bioturbation and presence of abundant fossils ofopen-marine invertebrates within pelleted mud (Figure2) indicate that this sediment was deposited in a shal-low subtidal setting in an environment similar toHolocene carbonate shelf and ramp settings.

Thin, generally less than 15 ft (4.5 m) thick, bio-herms composed of sponges, algae, corals, and bry-ozoans occur locally and are laterally discontinuous(cannot be correlated between wells) in the lower partof the open-marine section. Crinoid fragments are acommon accessory grain in this facies. Bioherms (Fig-ure 3), which are generally nonporous, contain abun-dant internal mud sediment that displays geopetalstructures. Skeletal grainstone, composed principallyof bryozoan and crinoid fragments and, less abun-dantly, fusulinid and mollusk fragments, is closelyassociated with bioherms. The presence of abundant

Figure 2. Core photograph of pellet packstone/grain-stone, which is a lithology in the open-marinefacies. The pellets originated as fecal pellets in a carbonate mud depositional environment. This rocktype is commonly porous and permeable; thisparticular sample also contains fusulinids [EastPenwell San Andres Unit No. 431, 3510 ft (1070 m),scale in centimeters].

Figure 3. Core photograph of a bioherm, which is alithology in the open-marine facies. Note the promi-nent bryozoan in growth position. Bioherms containinternal mud sediment and generally have lowporosity and permeability [Jordan University No.431, 3626 ft (1105 m), scale in centimeters].

Page 246: Reservoir Quality Prediction in Sand and Carbonates

fossils of open-marine organisms (Figure 4), lack ofdesiccation features, and stratigraphic proximity topellet packstone/grainstone indicate that biohermsand skeletal grainstone were deposited in a subtidal,open-marine environment.

Tidal-Flat Depositional Facies

Tidal-flat facies are pisolite packstone and mud-stone. Pisolite packstone is composed of poorly sortedsymmetrical and asymmetrical pisolites having diam-eters generally in the range of 0.2 to 4 mm andfine-grained carbonate mud matrix. Pisolites com-monly have a fitted fabric. This facies is characterizedby abundant caliche, fenestrae (Figure 5), desiccationcracks, tepee structures, and sheet cracks. Locally,minor karst dissolution is indicated by severe breccia-tion and infilling by greenish-gray siltstone. Thekarsted intervals are generally <3 ft (<1 m) thick. Thisfacies is commonly cemented with anhydrite andgypsum cement, and generally has very low porosityand permeability. Locally, however, cementation withcalcium sulfates is incomplete, and this facies may beporous and permeable. The abundant evidence forsyndepositional desiccation and the presence ofminor karst dissolution indicate that pisolite pack-stone formed in a tidal-flat environment that was fre-quently subaerially exposed.

Mudstone is composed of cream-colored, gener-ally massive dolomite, although some mudstone isfaintly laminated (Figure 6). Stromatolitic laminaeare present but rare. Mudstone is composed ofdolomite crystals generally smaller than 0.02 mm.With the exception of stromatolites, this facies is bar-ren of fossils, suggesting that it was deposited in ahypersaline environment in which stromatolites couldsurvive but marine invertebrates were excluded. The

absence of fossils and the close association with thepisolite packstone facies indicate deposition inhypersaline ponds on a tidal flat, isolated and proba-bly landward of an open-marine depositional envi-ronment.

Pisolite packstone and mudstone are interbeddedwith three intervals of siliciclastic silt that may becorrelated regionally using gamma-ray logs. Mostsiltstone is massive, although locally it is finely lami-nated. Siltstone is commonly calcareous and in tran-sitional contact with pisolite packstone andmudstone. The presence of siltstone interbeddedwith rocks containing evidence of subaerial exposure,and the lack of any regional sources for siliciclasticdetritus, suggest that these sediments were transportedto the tidal-flat environment by eolian processes. Somereworking in shallow water subsequent to eolian trans-port is indicated by the laminations.

Tidal-flat facies are separated from subjacent open-marine facies by an interval of greenish-grayorganic-rich shale that may be correlated throughoutJordan field using gamma-ray logs.

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 235

Figure 4. Core photograph of skeletal grainstone,part of the open-marine facies. This facies is closelyassociated with bioherms and is commonly porousand permeable. The circled feature is a well-preserved bryozoan [Jordan University No. 431, 3656 ft (1114 m), scale in centimeters].

Figure 5. Core photograph of pisolite packstone,which is a lithology in the tidal-flat facies. Thissample is porous because fenestrae are incompletelyfilled with sulfate cements [Jordan University No. 431, 3355 ft (1023 m), scale in centimeters].

Page 247: Reservoir Quality Prediction in Sand and Carbonates

236 Major and Holtz

Diagenetic Overprint

Tidal-flat pisolite packstone generally has low poros-ity because fenestrae and sheet cracks are cemented withcalcium sulfates. Locally, calcium sulfate cementationwas either incomplete, did not occur, or, more likely,sulfate cements were leached. Where little or no cementoccurs in pisolite packstone, this facies is porous andpermeable (Figure 5). The volumetrically dominant poretype is fenestral. This diagenetically controlled poroustexture is important because, where porous, the pisolitepackstone facies can be oil productive.

Open-marine facies have been partly to completelyaltered by a postburial leaching event. This diageneti-cally altered dolomite can be identified on slabbed coresurfaces as tan to brown rock that contrasts with thedark-gray color of unaltered dolomite. Altered dolomitein some cases mimics the geometry of burrows, whereasin other cases it forms aureoles around stylolites (Figure7), suggesting that the fluids causing this alteration pref-erentially flowed along stylolites (Carozzi and VonBergen, 1987; Von Bergen and Carozzi, 1990). This asso-ciation demonstrates that diagenetic alteration was apostburial, postcompaction event.

Diagenetically altered dolomite is more permeablethan unaltered dolomite, as indicated by miniperme-ameter data illustrated in Figure 8 (for description ofthis instrument, see Eijpe and Weber, 1971; Kittridge,1988; Chandler et al., 1989). The mottled geometry ofthis diagenetic alteration results in such close associa-tion of these two rock types that the order-of-magni-tude difference in permeability illustrated in Figure 8 iscommonly below the sampling resolution of conven-tional core-plug or whole-rock permeability analyses.

This permeability-enhancing diagenetic alteration isapparently the result of leaching and partial dissolu-tion of dolomite crystals. Hollow and corrodeddolomite crystals are visible at the light microscopeand scanning electron microscope level of resolution(Figure 9). Apparently this late-stage diagenetic eventwidened intercrystalline pore throats, resulting inincreased permeability. This diagenetic process mayalso partly cause alteration of nodules of anhydrite togypsum. Anhydrite nodules with outer edges alteredto gypsum are commonly surrounded by a “halo” of

Figure 6. Core photograph of mudstone, which is alithology in the tidal-flat facies. This lithology is near-ly completely nonporous [East Penwell San AndresUnit No. 431, 3329 ft (1015 m), scale in centimeters].

Figure 7. Pellet packstone/grainstone partiallyaltered by postburial leaching (area of lighter color).Note that the leached parts of this sample are associ-ated with stylolites, suggesting that undersaturatedfluids moved along stylolites [Jordan University No.638W, 3478 ft (1060 m), core is ~3 in. (7.7 cm) wide].

Page 248: Reservoir Quality Prediction in Sand and Carbonates

altered dolomite (Figure 10), and some samples of dia-genetically altered rock contain ≤20% gypsum. It can beinferred from this relationship that the fluids leachingthe dolomite also altered (hydrated) some of the anhy-drite nodules and cements to gypsum. Inasmuch asgypsum in San Andres Formation and Grayburg For-mation reservoirs is restricted to the central and south-ern parts of the Central Basin Platform, and formationwater resistivities in this area increase in a southerlydirection (M.H. Holtz and R.P. Major, 1995, unpub-lished data), we infer that the fluids that created thehigh-permeability diagenetically altered dolomite atJordan field originated from the south.

The unaltered and altered dolomite textures havesimilar carbon isotope compositions, but they may bedistinguished by different ranges of oxygen isotopecomposition (Figure 11). Carbon isotopic compositionsare generally in the range of 4.5 to 6‰. The unaltered

dolomites have oxygen isotope compositions thatrange from 3 to 5.5‰, whereas the altered dolomitehas more depleted oxygen isotope compositions of1–4‰.

A similar range of isotopic compositions in SanAndres dolomites from the Central Basin Platformwas interpreted by Leary and Vogt (1990) as indicat-ing that the altered dolomite was recrystallizedeither at elevated temperatures or in the presence ofwater with a depleted oxygen isotope composition.As outlined above, however, the textures of thealtered dolomite suggest an episode of leaching. Forexample, textural evidence suggests that diageneticalteration included removal of some cloudy, inclu-sion-rich cores of dolomite rhombs.

Dolomitization of the San Andres Formation hasbeen attributed to hypersaline fluids on the basis ofstratigraphic proximity to overlying evaporites anddolomite geochemistry (for example, Bein and Land,1983). If the cloudy dolomite cores in the Jordan SanAndres reservoir were formed from hypersalinebrines early in the diagenetic history of these rocks,it can be inferred that the inclusion-rich cores ofthese crystals are enriched in 18O relative to thelimpid rims of the crystals. Thus, textural as well asgeochemical data suggest that the diageneticallyaltered texture is the result of a permeability-increasing leaching event that preferentially removed18O-enriched, presumably less stoichiometric,dolomite. This resulted in a bulk rock with relativelydepleted 18O composition and enhanced permeabil-ity. Alternatively, the light oxygen isotope signatureof the altered texture may be the result of recrystal-lization that preceded leaching; the limpid rims ofaltered dolomite rhombs have a different oxygen iso-tope composition than do the rims of unaltereddolomite rhombs. This cannot be determined withcurrently available sampling technology for isotopeanalysis.

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 237

Figure 8. Permeability histograms for tidal-flat facies(pisolite packstone), unaltered marine facies (pelletpackstone/grainstone), and altered marine facies (pellet packstone/grainstone). The highestpermeabilities are in diagenetically altered pelletpackstone/grainstone. Data are from West JordanUnit No. 12-4, East Penwell San Andres Unit No. 207,and East Penwell San Andres Unit No. 1914. All datawere collected from cores that were not subjected tohigh-temperature, gypsum-destructive handling.

Figure 9. Thin-section photomicrograph illustratingleached, hollow dolomite rhombs in diageneticallyaltered pellet packstone/grainstone. [East PenwellSan Andres Unit No. 1313, 3922 ft (1495 m), dolomitecrystals are ~50 mm in width].

Page 249: Reservoir Quality Prediction in Sand and Carbonates

238 Major and Holtz

Flow Units

Because responses of neutron and acoustic logs togypsum-bearing rocks differ, these two logs can beused to identify diagenetically altered rock textures inwells that are not cored. As indicated previously, thehigh-permeability diagenetically altered rock is associ-ated with higher gypsum content than unaltered rock.Thus, altered reservoir rock containing abundant gyp-sum may be identified on wireline logs wheredolomitic neutron log porosity exceeds acoustic poros-ity normalized to a dolomite matrix. The relationshipof acoustic log porosity, neutron porosity, and percentof altered texture observed in slabbed core demon-strates the use of wireline logs to identify the diagenet-ically altered facies (Figure 12).

Jordan San Andres reservoir is divided into fourflow units (sensu Ebanks, 1987) on the basis of bothdepositional facies and diagenetic overprint. Open-marine rocks are divided into three flow units definedby the stratigraphic patterns of diagenetically altered

facies as identified using wireline logs. The lowermostflow unit A is 100%, or nearly 100%, altered-texture rockand is characterized by a neutron log–acoustic logporosity-curve separation. The overlying flow unit B iscomposed of diagenetically unaltered rock character-ized by a normalized neutron log that is in good agree-ment with a normalized acoustic log. The overlying flowunit C is composed of a mottled mixture of diageneti-cally altered and unaltered rock and is characterized bya neutron log–acoustic log separation (Figures 12, 13).

The uppermost flow unit D is composed of tidal-flatrocks that occur above the organic-rich shale identifiedby a gamma-ray marker (Figure 13). This marker can becorrelated across the field. Porosity in this section occursin pisolite packstone in which fenestrae and sheet cracksare not plugged with calcium sulfate cements.

Open-marine rocks contain thin zones of siliciclasticsilt concentrated along stylolite swarms and shalypartings, many of which can be correlated throughoutthe field using gamma-ray logs. These intervals areinterpreted to document low sedimentation rates asso-ciated with rapid rise of relative sea level. As such,these beds define cycle boundaries even where overly-ing and underlying depositional textures are nearlyindistinguishable. Indeed, correlation of these featureswestward (updip) indicates that in more landwarddepositional environments these surfaces dividecycles defined by upward-shallowing depositional

Figure 10. Core photograph of anhydrite nodulespartly altered to gypsum in diagenetically alteredpellet packstone/grainstone. Note that the diageneti-cally altered (lighter colored) pellet packstone/grainstone is associated with alteration of gypsumto anhydrite [East Penwell San Andres Unit No.1914, 3918 ft (1194 m)].

Figure 11. Carbon and oxygen isotope cross plot ofdiagenetically altered and unaltered pellet pack-stone/grainstone. The unaltered rock has oxygen iso-tope compositions of 3‰–5.5‰, and the altered rockhas compositions of 1‰–4‰. All stable isotope datareported relative to the PDB standard.

Page 250: Reservoir Quality Prediction in Sand and Carbonates

facies (S.C. Ruppel, 1995, personal communication).Thus, these gamma-ray correlations are approximatetime lines (Ruppel et al., 1995).

The boundaries of between flow units A and B andflow units B and C are approximately parallel to time

lines, which suggests that the shaly partings and stylo-lite swarms acted as aquatards to flow of the fluidsthat leached the open-marine parts of this reservoir.Note, however, that flow unit C pinches out in adowndip direction. Thus, despite the influence ofdepositional textures on flow-unit geometry, diage-netic overprint crosscuts depositional facies and is themajor control on flow-unit geometry (Figure 13).

PETROPHYSICAL RESERVOIRDESCRIPTION

The second step in predicting reservoir quality is tomake a petrophysical description of the reservoir. Inthis step, we take the geologic reservoir model, inter-preted in terms of flow units, and describe porosity,permeability, and estimates of initial water and oil sat-uration within that geologic context.

Calibration of Porosity Between Logs and Cores

Although acoustic, neutron, and density porositylogs are available at Jordan field, the presence of abun-dant gypsum in this reservoir precludes the use ofneutron and density logs for reliable porosity mea-surements. Neutron logs measure the bound water ofhydration in gypsum as porosity, and the low densityof gypsum (2.35 g/cm3) relative to that of dolomite(2.88 g/cm3) and anhydrite (2.98 g/cm3) results in sig-nificant uncertainty concerning matrix density used indensity log calculations (Tilly et al., 1982; Bebout et al.,1987; Holtz and Major, 1994). Furthermore, calibrationof logs with core data necessitates the use of coreporosity data collected using low-temperaturenongypsum-destructive analytic techniques. Conven-tional high-temperature analysis volatilizes the boundwater of hydration in gypsum crystals and yieldsincorrect values.

The most reliable log-derived porosities were madefrom a calibration of acoustic transit time with coreporosity measured by low-temperature, nondestruc-tive techniques. Use of all three open-hole porositytools resulted in a poorer fit than the use of acousticlogs alone. The Jordan field porosity-acoustic transittime relationship is

(1)

where φ= porosity (%) and ∆t = acoustic transit time ortwo-way traveltime (µsec/ft).

Permeability Character

Division of the reservoir into flow units is based onidentification of three rock types with distinctly differ-ent permeabilities: tidal-flat facies, unaltered open-marine facies, and altered open-marine facies.Minipermeameter data are used to evaluate open-marine facies rocks because available low-temperaturewhole-core data do not have sufficient sampling reso-lution to discriminate between unaltered and alteredopen-marine facies (there are no low-temperature coreanalyses from flow unit A). The geometric mean of

φ = +– . .44 159 1 006∆t

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 239

Figure 12. Calibration of acoustic log porosity, neu-tron log porosity, and percent of diageneticallyaltered facies observed in core. Note that the flowunits and depositional rock types do not correlate.Flow units are dominantly controlled by diageneticalteration. Data from Jordan University No. 114.

Page 251: Reservoir Quality Prediction in Sand and Carbonates

240 Major and Holtz

Figu

re 1

3. W

est-

east

cro

ss s

ecti

on th

rou

gh J

ord

an (S

an A

nd

res)

fie

ld il

lust

rati

ng

flow

un

its.

Dia

gen

etic

ally

alt

ered

, hig

h-p

erm

eab

ilit

y zo

nes

(sh

aded

) are

rea

dil

y d

isti

ngu

ish

ed b

y co

mp

arin

g n

eutr

on a

nd

aco

ust

ic lo

g p

oros

ity,

bec

ause

the

latt

er is

aff

ecte

d b

y th

e p

rese

nce

of

gyp

sum

.O

RS

= o

rgan

ic-r

ich

sh

ale.

Page 252: Reservoir Quality Prediction in Sand and Carbonates

minipermeameter data is 1.3 md for the unalteredopen-marine facies and 8.4 md for altered open-marinefacies (Figure 8). A relatively small amount of low-temperature, conventional core permeability data isavailable from tidal-flat facies cored in wells offsettingUniversity Lands. The geometric mean of these data is2.1 md (Figure 8), although the variance of these data ismuch greater than that of other rock types due to thelarge percentage of vuggy porosity in this rock type.

Core plugs carefully chosen to sample flow unitrock types were analyzed for porosity and permeabil-ity. Porosity-permeability relationships for the threerock types demonstrate that, for any porosity >10%,tidal-flat facies rocks have the highest permeabilities,followed by altered open-marine rocks and unalteredopen-marine rocks. For porosities <10%, tidal-flatrocks have only slightly higher permeabilities thanunaltered open-marine rocks (Figure 14).

Empirical equations that describe the relationshipbetween porosity and permeability were derived foreach flow unit. Because the relationship between poros-ity and permeability for tidal-flat rocks is bimodal, thereare separate equations for flow unit D rocks having<10% porosity and >10% porosity. The 10% porositythreshold corresponds to the point at which vuggyporosity becomes interconnected, causing higher per-meability per incremental increase in porosity.

(2)

(3)

(4)

(5)

(6)

where k = permeability (md) and φ= porosity (%).The geometric mean of vertical permeability in West

Jordan Unit Well No. 17-2 is 0.15 md and that of hori-zontal permeability is 0.3 md [similar to the relation-ship for the San Andres reservoir at Slaughter field(Ebanks, 1990)]. This suggests that vertical flow canoccur on a small scale but not on a large scale. How-ever, the Dystra Parsons coefficient value of 0.93 inwell 17-2 documents a high degree of vertical perme-ability variation. This suggests that, on a 1-ft (0.3-m)scale, oil can flow vertically if it can flow horizontally.On a larger scale of tens of feet, permeability varieswidely. Indeed, Ford and Kelldorf (1976) demonstratedthat zones of bypassed oil and high-permeability “thiefzones” occur in the Jordan (San Andres) reservoirbecause of the lack of vertical permeability continuity.

Capillary Pressure Curves

Nearly all of the resistivity logs in the Jordan (SanAndres) reservoir postdate initiation of the waterflood.Thus, the formation waters at the time these logs weremade may have been contaminated with floodwaters,and some zones may have been more thoroughlyflushed than others. Under these circumstances, water

resistivity is unknown; therefore, resistivity data cannotbe used to calculate original saturations. An alternateapproach is to use capillary-pressure data to estimatehydrocarbon saturation, following the procedure ofAmyx et al. (1960). Nine plug samples were selectedfrom undamaged parts of cores that had been previ-ously sampled for high-temperature core analysis.These were analyzed by brine-injection capillary-pressure tests, using procedures that do not dehydrategypsum. Plugs were chosen to represent the three prin-cipal rock types: diagenetically unaltered open-marinefacies, diagenetically altered open-marine facies, andporous tidal-flat facies.

The shapes of capillary-pressure curves indicatethe nature of the pore structure (Murray, 1960; Keithand Pittman, 1983). Both altered and unaltered open-marine rocks (Figure 15a) have capillary-pressurecurve shapes that suggest unimodal pore-throat sizedistribution, thus substantiating the observation thateither interparticle or intercrystalline pores domi-nate. This indicates that vugs do not make a majorcontribution to porosity, and explains why the coreporosity–acoustic transit time function is a reliablemethod of calculating log-derived porosity. Theshape of tidal-flat facies capillary-pressure curves(Figure 15b) also indicates a unimodal pore-throat-size distribution. However, the variability in pore-throat size is less for tidal-flat facies than for alteredand unaltered open-marine facies. This suggests thatthe fenestral pores are not connected as touchingvugs but, rather, are connected by sheet cracks ofsimilar pore-throat size.

Capillary-pressure curves may be used to estimateoriginal water and hydrocarbon saturations in variousrock types provided that permeability and heightabove the free-water table are known (Amyx et al.,1960). The height of the free-water table in the Jordan(San Andres) reservoir is estimated to be 950 ft (290 m)subsea, as will be discussed in the next section. Analy-sis of capillary-pressure data by applying multiplenonlinear regression resulted in two equations, one

Flow unit D 10% porosity: ( ) : . * .

>= × −k 4 096 10 103 0 329φ

Flow unit D 10% porosity: ( ) : . * .

<=k 0 0271 100 125φ

Flow unit C : . * .k = × −3 08 10 103 0 209φ

Flow unit B : . * .k = × −2 5 10 103 0 188φ Flow unit A : . * .k = × −3 5 10 103 0 234φ

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 241

Figure 14. Relationship between porosity and per-meability for the unaltered open-marine facies,altered open-marine facies, and tidal-flat facies.

Page 253: Reservoir Quality Prediction in Sand and Carbonates

242 Major and Holtz

that calculates water saturation for altered and unal-tered open-marine facies (flow units A, B, and C), andone that calculates water saturation for tidal-flat faciesfacies (flow unit D).

(7)

(8)

where Swi = initial water saturation (%) and haw =height above oil-water contact (ft).

Equation 7 is statistically significant, having an Fvalue (analysis of variance) of 80.4 (F critical = 3.15) at a95% confidence level and an r2 of 0.84. Equation 8 is alsostatistically significant at a 95% confidence level with anF value of 60.69 (F critical = 3.49) and an r2 of 0.87.

RESOURCE EVALUATION

Interpretation of reservoir volumes is the last stepin a comprehensive prediction of reservoir quality.The goal of this part of the study is to quantitativelyestimate the amount of remaining oil and to map itsdistribution in the reservoir.

Production Patterns

Per-well production data are available only for thepostwaterflood time period, 1969 to the present. Pro-duction data before 1969 are available on a per-leasebasis (the two Jordan San Andres units on UniversityLands were composed of several separate leases dur-ing different periods). Flow tests were performed peri-odically on individual wells, and these test data wereused to apportion annual lease production to eachwell within the lease.

The updip part of the reservoir has been moredensely drilled than other parts (Figure 1). To compen-sate for variations in well density, the drainage area ofeach well was approximated within a grid of square40-ac cells. The fraction of well drainage areas withineach 40-ac cell was used to apportion production.Thus, a single data point for each 40-ac cell expressesproduction in units of million stock tank barrels peracre. These data were contoured to produce a cumula-tive production map (Figure 16). This map exhibits atrend of high production extending from the updipcentral-western margin of the field to the downdipsoutheastern corner. The southwestern corner of thefield is an area of low production.

Original Oil in Place

Calculation of original oil in place (OOIP) requiresknowledge of the height above free-water level. Welllogs cannot be used to calculate water saturation, andthus identify the depth at which water saturations are100%, because all resistivity logs in the Jordan (SanAndres) reservoir postdate waterflooding, and origi-nal water resistivities are unknown. Two observationswere used to estimate the elevation of free water. First,core analysis data indicate no hydrocarbon saturationbelow 950 ft (290 m) subsea. A core from UniversityWell No. 638W contained no residual oil saturationbelow 950 ft (290 m) subsea; a core from University WellNo. 114 had residual oil saturation to 930 ft (283 m) sub-sea, at which depth the rock has less than 0.1 md per-meability; and a core from University Well No. 219Whad no residual oil saturation below 945 ft (288 m) sub-sea. Second, shallow-resistivity values recorded bymicroresistivity logs were compared with medium- anddeep-resistivity values in porous zones. These resistivi-ties indicate that, below 950 ft (290 m) subsea, there was

Flow unit D :. . * log( ) . * log( )S h kwi aw= − −68 5 19 7 6 59

Flow units A, B, and C :. . * log( ) . * log( )S h kwi aw= − −120 22 36 016 11 968

Figure 15. Brine capillary-pressure curves for the unaltered and altered open-marine facies (a) andtidal-flat facies (b).

Page 254: Reservoir Quality Prediction in Sand and Carbonates

no flushing of movable oil adjacent to well bores.Thus, we interpret the free-water level to be 950 ft(290 m) subsea. This interpretation is supported by theresults of well-deepening associated with initiation ofwaterflooding in the 1970s. Wells that did not previ-ously reach the 9-ft (2.7-m) free-water level and weredeepened [generally to depths ≤950 ft (290 m)] beganproducing oil at greater rates, indicating that much ofthe lower part of the reservoir had not been drained.

Many wells in the Jordan (San Andres) reservoir donot penetrate the lower part of the reservoir, eventhough some wells were deepened in the 1970s. Forthis reason, we have divided the reservoir into twozones for the purpose of evaluating remaining oil. Amap of total depths of wells (Figure 17) illustrates thegeometry of the part of the reservoir between totaldepths of wells and the free-water level of 950 ft(290 m) subsea. Remaining oil in the reservoir was cal-culated separately for the contacted and uncontactedparts of the reservoir using an oil formation volumefactor of 1.28 (Holtz et al., 1991).

Contacted Reservoir

Original oil in place was mapped in units of oil sat-uration ×porosity ×thickness (Soφh) for each flow unit.Saturations were calculated using equations 7 and 8with porosity values from acoustic logs calculatedusing equation 1. Volumes of original oil in place werecalculated by gridding Soφh values over their represen-tative areas and multiplying Soφh by area to yield vol-ume. This yielded original oil-in-place values of 36.3

MMSTB for flow unit A, 53.5 MMSTB for flow unit B,30.9 MMSTB for flow unit C, and 96.8 MMSTB for flowunit D (Figure 18). The sum of original oil in place cal-culated for all four flow units is 218 MMSTB, whichyields a 31.2% recovery efficiency.

Maps of Soφh illustrate the spatial distribution oforiginal oil in place for each flow unit. The flow unit Amap (Figure 18a) indicates that the highest oil volumeswere contained in the updip central part of the field.Flow unit A is close to the free-water level in thedowndip (east) part of the field; the zero contourmarks the point at which the top of flow unit A meetsthe free-water level. The flow unit B Soφh map (Figure18b) indicates highest oil volumes in the updip centralpart of the field and in the southwest part of the field.The volume of oil generally decreases downdip. Flowunit B is absent in the west-central and southeast partsof the field; therefore, flow unit C is immediatelysuperjacent to flow unit A in these areas. The flow unitC map (Figure 18c) illustrates an area of highest oilvolume crosscutting structure from the downdipsoutheastern part of the field to the updip central partof the field. Flow unit C is absent in the northeast andsouthwest areas. The flow unit D map (Figure 18d)indicates generally increasing oil volumes in thedowndip (east) part of the field and in the updipnorth-central part of the field.

Uncontacted Reservoir

Estimating the amount of original oil in place inuncontacted parts of the reservoir clearly requiressome assumptions. We start with the conservativeassumption that those parts of the reservoir belowtotal depths of wells will, in most instances, be close to

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 243

Figure 16. Cumulative production map based onreconstructed per-well production, 1937–1988, nor-malized on a 40-ac grid. Note that the axis of highestcumulative production has a northwest-southeastorientation. MSTB/ac = million stock tank barrelsper acre.

Figure 17. Contour map of total depths of wells. Notethat most wells do not reach 950 ft (290 m) subsea,which is the interpreted free-water level.

Page 255: Reservoir Quality Prediction in Sand and Carbonates

244 Major and Holtz

Figure 18. (a) Flow unit A oil saturation ×porosity ×thickness Soφh map. The zero contour on the western(downdip) side of the field indicates where flow unit A dips below the free-water level. (b) Flow unit B Soφhmap. (c) Flow unit C Soφh map. The zero contours indicate that this flow unit pinches out to the northeast andsouthwest. (d) Flow unit D Soφh map.

Page 256: Reservoir Quality Prediction in Sand and Carbonates

the free-water level and, therefore, have relatively lowsaturations because of buoyancy considerations(Amyx et al., 1960). In the absence of log data, we esti-mated original oil in place by calculating the averageSoφh/gross pay ratio, by flow unit, where well-log dataare available and, importantly, where the flow unit isnear the free-water level. This ensures that the databasewe use for estimating the average Soφh/gross pay ratiocontains only those parts of each flow unit that haverelatively low oil saturations. The average Soφh/grosspay ratios for flow units A and B are illustrated in Fig-ure 19. There were only two logged wells that con-tained flow unit C and one that contained flow unit Din contact with free water; we have made calculationsfor these flow units using the flow unit B average fromFigure 19 because this is the lowest value and we seekto make conservative estimates.

The uncontacted part of the reservoir contains anOOIP of 23 MMSTB: 6.6 MMSTB in flow unit A, 4.7MMSTB in flow unit B, 1.0 MMSTB in flow unit C, and10.7 MMSTB in flow unit D. The highest volumes ofuncontacted oil are in the southwest and southeastcorners of the field. Some local high values are presentin the central part of the field (Figure 20).

Distribution of Remaining Oil

Our calculation of total OOIP for both contacted anduncontacted parts of the reservoir is 240.5 MMSTB,which indicates a recovery efficiency of 28%. Thus, 173MMSTB remain in this reservoir. By applying an aver-age saturation of oil residual to waterflood for the SanAndres/Grayburg reservoirs (Finley et al., 1990), weestimate that 113 MMSTB are residual oil and 59.5MMSTB of remaining oil are mobile to waterflood.

Contacted Reservoir

Within the contacted part of the reservoir thereremain 47.3 MMSTB of mobile oil and 102.2 MMSTB ofoil residual to waterflood. To obtain a conservative

remaining oil distribution, all oil production wasassumed to have come from the contacted part of thereservoir. Cumulative production (Figure 16) was sub-tracted from OOIP, calculated by summing OOIP in allflow units (Figure 18), resulting in a remaining-oilmap for the contacted zone (Figure 21).

Four areas of high remaining oil occur in the con-tacted part of the reservoir (Figure 21). The north-trending area of high remaining oil in the easterncentral part of the field is coincident with a relativehigh OOIP in flow unit D (Figure 18d), suggesting thatmost of this oil is in flow unit D. The high remainingoil in the updip western-central part of the field isapproximately coincident with high OOIP values in allflow units (Figure 18). The area of high remaining oilin the northern part of the field is approximately coin-cident with high OOIP in flow unit D and, to a lesserextent, flow unit B, and the high remaining oil in thesouthwest corner of the field is also approximatelycoincident with high OOIP in flow units B and D (Fig-ure 18b, d). This suggests that this oil is primarily inflow units B and D. Most of the remaining oil in thecontacted part of this reservoir is in low-permeability,diagenetically unaltered marine facies of flow unit Band tidal-flat facies of flow unit D.

Because the remaining-oil map (Figure 21) illus-trates volumes of remaining mobile oil and residual oil,it is difficult to evaluate the extent to which thisremaining resource can be produced by increasedsweep efficiency. Unfortunately, there are no irre-ducible oil saturation values for each of the flow unitsidentified in the Jordan (San Andres) reservoir. How-ever, using an average irreducible oil saturation valueof 32% for dolomitized San Andres and Grayburgreservoirs in the Central Basin Platform (Finley et al.,

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 245

Figure 19. Cumulative frequency of net Soφh/grosspay ratio for flow units A and B for wells in whichthese flow units are in contact with free water.

Figure 20. Contour map of Soφh for the uncontactedpart of the reservoir.

Page 257: Reservoir Quality Prediction in Sand and Carbonates

246 Major and Holtz

1990), we can calculate the approximate sweep effi-ciency based on cumulative production. A map of cur-rent recovery efficiency, calculated as percent oforiginal mobile oil in place that has been produced,indicates that the north, east, and southwest parts ofthe field have low recovery efficiencies (Figure 22).

Uncontacted ReservoirAbout 12.2 MMSTB of mobile oil and 10.8 MMSTB

of residual oil remain within the uncontacted part ofthe reservoir, based on the assumption that cumula-tive production from the field has come exclusivelyfrom the contacted part of the reservoir. A large area ofhigh remaining oil occurs in the southeast corner ofthe field (Figure 20), where flow unit A is below free-water level (Figure 18a) and flow unit B is absent (Fig-ure 18b), indicating that most of the oil in this area ofthe uncontacted part of the reservoir is in flow units Cand D. A large area of high remaining oil also occurs inthe southwest corner of the field, where most wells donot penetrate below 800 ft (244 m) subsea. In this area,flow unit A is below the oil-water contact (Figure 18a)and flow unit C is absent (Figure 18c), indicating thatmost of the oil in this area is in flow units B and D.

APPLICATION OF RESERVOIR QUALITY PREDICTION

This prediction of reservoir quality in the Jordan(San Andres) field provides three avenues forincreased production: (1) focusing the waterflood inselected areas of the contacted part of the reservoir,(2) deepening wells to reach the uncontacted part of thereservoir, and (3) initiation of a carbon dioxide flood.

The parts of the contacted reservoir that are mostprospective for waterflood profile modification occur inflow unit B, which is characterized by low permeability,and flow unit D, which is characterized by highly het-erogeneous permeability. The largest area of unsweptremaining mobile oil in flow unit B is in the southwestcorner of the field; the largest area of unswept mobileoil in flow unit D is in the east-central part of the field(Figures 21, 22). In these areas, selective well-bore plug-ging and perforation squeezing could focus the water-flood and increase ultimate oil recovery.

Large parts of the reservoir are below total depthsof wells and, therefore, have not been drained. Con-sider the high cumulative production trend thatextends from the updip central-western part of thefield to the downdip southeastern corner (Figure 16)and the somewhat similar trend of areas in which wellbores have penetrated to the free-water depth of 950 ft(290 m) subsea (Figure 17). The areas of highest cumu-lative production are due, at least in part, to well boreshaving been drilled through the entire pay zone. Themost prospective areas for deepening wells are thosein which high-permeability flow units A and C areuncontacted. The largest area of uncontacted flow unitC is in the southeast corner of the field. Most of flowunit A that is above the free-water level is at least par-tially penetrated by well bores, but there are deepen-ing opportunities in this flow unit in the central part ofthe field and near the central-western margin of thefield. There are large volumes of remaining oil in flowunits B and D in the southwest corner, and flow unit Din the southeast corner, although efficient productionfrom these areas will probably require focusing thewaterflood. Because porosity and permeability in

Figure 21. Contour map of remaining oil in the con-tacted part of the reservoir, constructed by subtract-ing cumulative production from original oil inplace. MSTB = thousand stock tank barrels.

Figure 22. Contour map of sweep efficiency definedas the percent of original mobile oil in place that hasbeen produced.

Page 258: Reservoir Quality Prediction in Sand and Carbonates

tidal-flat facies are extremely heterogeneous and pre-diction is unreliable, we assign a low priority to welldeepening in the southeast corner of the field.

The Jordan (San Andres) reservoir is a good candi-date for increased ultimate recovery by miscible-gasenhanced oil recovery operations. The reservoir depth,oil gravity, current oil saturation, and oil viscosity sug-gest that this reservoir is an excellent candidate for car-bon dioxide flooding, according to the criteria ofStalkup (1983). Taber and Martin (1992) predicted thatcarbon dioxide flooding would increase averagerecovery for West Texas San Andres reservoirs by 11%of OOIP. Application of this prediction to the Jordan(San Andres) reservoir indicates a potential for anadditional 24 MMSTB of oil production from the con-tacted part of the reservoir.

CONCLUSIONS

Reservoir quality in Jordan (San Andres) field is pre-dicted for four flow units identified on the basis of bothdepositional facies and subsequent diagenetic alteration.Volumetric calculations and cumulative production pat-terns indicate that of an original 240 MMSTB of OOIP,113 MMSTB of which are residual and 127.5 MMSTB ofwhich are mobile, ~59.5 MMSTB of mobile oil remain asboth bypassed and uncontacted oil. The largest vol-umes of bypassed oil occur in low-permeability flowunit B and heterogeneous-permeability flow unit D inareas of low cumulative production. Waterflood profilemodification by selective perforation squeezing mayfocus injection water into the flow units in these areasand contact bypassed oil that would otherwise remainunrecovered. Many of the wells in this reservoir do notpenetrate to the free-water level; parts of the reservoirin which the high-permeability flow units A and C areuncontacted by well bores are the principal targets forincreased production by well-bore deepening. The Jor-dan (San Andres) reservoir has physical characteristicsthat make it an excellent candidate for enhanced oilrecovery by carbon dioxide flood.

ACKNOWLEDGMENTS

Funding for this study was provided by The Univer-sity of Texas System as part of a larger study of reser-voirs on The University of Texas Lands. Shell OilCompany, Hondo Oil Company, The University ofTexas Lands Office, and the Railroad Commission ofTexas provided access to data. We thank M.G. Kittridgefor minipermeameter data that were collected in thelaboratories of the Department of Petroleum andGeosystems Engineering, The University of Texas atAustin. J.E. Nicol and Mohammed Sattar providedtechnical support. Brine capillary-pressure analyseswere conducted by Bell Laboratories (Midland), andRadian Corporation donated mapping software. We aregrateful for the review comments of W.A. Ambrose,P.M. Harris, T.F. Hentz, J.A. Kupecz, and W.G. Zem-polich. Parts of this paper have been published by theSociety of Petroleum Engineers in the Journal of

Petroleum Technology (1990, v. 42, no. 10, p. 1304–1309)and Permian Basin Oil and Gas Recovery Conference Pro-ceedings (1994, p. 565–576). This material is used herewith permission of the society. Publication authorizedby the director of the Bureau of Economic Geology, TheUniversity of Texas at Austin.

REFERENCES CITED

Amyx, J.W., D.M. Bass, Jr., and R.L. Whiting, 1960,Petroleum Reservoir Engineering: New York,McGraw Hill Book Co., 610 p.

Bebout, D.G., F.J. Lucia, C.R. Hocott, G.E. Fogg, andG.W. Vander Stoep, 1987, Characterization of theGrayburg reservoir, University Lands Dune Field,Crane County, Texas: The University of Texas atAustin, Bureau of Economic Geology Report ofInvestigations No. 168, 98 p.

Bein, A., and L.S. Land, 1983, Carbonate sedimentationand diagenesis associated with Mg-Ca-Chloridebrines: the Permian San Andres Formation in theTexas Panhandle: Journal of Sedimentary Petrology,v. 53, p. 243–260.

Carozzi, A.V., and D. Von Bergen, 1987, Styloliticporosity in carbonates: a critical factor for deephydrocarbon production: Journal of PetroleumGeology, v. 10, p. 267–282

Chandler, M.A., D.J. Goggin, and L.W. Lake, 1989, Amechanical field permeameter for making rapid, non-destructive permeability measurements: Journal ofSedimentary Petrology, v. 59, p. 613–615.

Ebanks, W.J., Jr., 1987, Flow unit concept—integratedapproach to reservoir description for engineeringprojects (abs.): AAPG Bulletin, v. 71, p. 551–552.

Ebanks, W.J., Jr., 1990, Geology of the San Andresreservoir, Mallet Lease, Slaughter field, HockleyCounty, Texas: implications for reservoir engineer-ing projects, in D.G. Bebout and P.M. Harris, eds.,Geologic and engineering approaches in evaluationof San Andres/Grayburg hydrocarbon reservoirs—Permian Basin: The University of Texas at Austin,Bureau of Economic Geology, p. 75–85.

Eijpe, R., and K.J. Weber, 1971, Mini-permeameters forconsolidated rock and unconsolidated sand: AAPGBulletin, v. 55, p. 307–309.

Finley, R.J., S.E. Laubach, N. Tyler, and M.H. Holtz,1990, Opportunities for horizontal drilling inTexas: The University of Texas at Austin, Bureau ofEconomic Geology Geological Circular 90-2, 32 p.

Ford, W.O., Jr., and W.F.N. Kelldorf, 1976, Field resultsof a short-setting-time polymer placement technique:Journal of Petroleum Technology, v. 28, p. 749–756.

Galley, J.E., 1958, Oil and geology in the Permian Basinof Texas and New Mexico, in L.G. Weeks, ed., Habitatof oil: AAPG, Tulsa, OK, p. 395–446.

Harris, P.M., C.A. Dodman, and D.M. Bliefnick, 1984,Permian (Guadalupian) reservoir facies, McElroyfield, West Texas, in P.M. Harris, ed., Carbonatesands—a core workshop: SEPM Core Workshop 5,p. 136–174.

Harris, P.M., and S.D. Walker, 1990, McElroy field:Development geology of a dolostone reservoir, Per-mian Basin, West Texas, in D.G. Bebout and

Predicting Reservoir Quality at the Development Scale: Methods for Quantifying Remaining Hydrocarbon Resource 247

Page 259: Reservoir Quality Prediction in Sand and Carbonates

248 Major and Holtz

P.M. Harris, Geologic and engineering approaches inevaluation of San Andres/Grayburg hydrocarbonreservoirs—Permian Basin: The University of Texasat Austin, Bureau of Economic Geology, p. 275–296.

Holtz, M.H., and R.P. Major, 1994, Effects of deposi-tional facies and diagenesis on calculating petro-physical properties from wireline logs in Permiancarbonate reservoirs of West Texas (abs.): AAPGBulletin, v. 78, p. 494–495.

Holtz, M.H., N. Tyler, C.M. Garrett, Jr., W.G. White,and N.S. Banta, 1991, Atlas of major Texas oil reser-voirs database: The University of Texas at Austin,Bureau of Economic Geology, 1 disk.

Keith, B.D., and E.D. Pittman, 1983, Bimodal porosity inoolitic reservoirs—effect on productivity and logresponse, Rhodessa Limestone (Lower Cretaceous),East Texas Basin: AAPG Bulletin, v. 67, p. 1391–1399.

Kerans, C., F.J. Lucia, and R.K. Senger, 1994, Inte-grated characterization of carbonate ramp reser-voirs using Permian San Andres Formation outcropanalogs: AAPG Bulletin, v. 78, p. 191–216.

Kittridge, M.G., 1988, Analysis of areal permeabilityvariations—San Andres Formation (Guadalupian):Algerita escarpment, Otero County, New Mexico:M.S. thesis, The University of Texas, Austin, Texas,361 p.

Leary, D.A., and J.N. Vogt, 1990, Diagenesis of the SanAndres Formation (Guadalupian), Central BasinPlatform, West Texas, in D.G. Bebout and P.M. Har-ris, eds., Geological and engineering approaches inevaluation of San Andres/Grayburg hydrocarbonreservoirs—Permian Basin: The University of Texasat Austin, Bureau of Economic Geology, p. 21–47.

Longacre, S.A., 1980, Dolomite reservoirs from Per-mian biomicrites, in R.B. Halley and R.G. Loucks,eds., Carbonate reservoir rocks: SEPM Core Work-shop 1, p. 105–117.

Longacre, S.A., 1983, A subsurface example of adolomitized middle Guadalupian (Permian) reeffrom West Texas, in P.M. Harris, ed., Carbonatebuildups—a core workshop: SEPM Core Workshop4, p. 304–326.

Major, R.P., D.G. Bebout, and F.J. Lucia, 1988, Deposi-tional facies and porosity distribution, Permian(Guadalupian) San Andres and Grayburg forma-tions, PJWDM field complex, Central Basin Plat-form, West Texas, in A.J. Lomando and P.M. Harris,eds., Giant oil and gas fields: SEPM Core Workshop12, p. 615–648.

Major, R.P., G.W. Vander Stoep, and M.H. Holtz, 1990,Delineation of unrecovered mobile oil in a maturedolomite reservoir: East Penwell San Andres Unit,University Lands, West Texas: The University ofTexas at Austin, Bureau of Economic GeologyReport of Investigations No. 194, 56 p.

Murray, R.C., 1960, Origin of porosity in carbonaterocks: Journal of Sedimentary Petrology, v. 30, p. 59–84.

Ruppel, S.C., C. Kerans, R.P. Major, and M.H. Holtz,1995, Controls on reservoir heterogeneity in Per-mian shallow-water platform carbonate reservoirs,Permian Basin: implications for improved recovery:The University of Texas at Austin, Bureau of Eco-nomic Geology Geologic Circular 95-2, 30 p.

Stalkup, F.I., Jr., 1983, Miscible displacement: Societyof Petroleum Engineers Monograph Series 8, 204 p.

Taber, J.J., and D.F. Martin, 1992, Carbon dioxideflooding: Journal of Petroleum Technology, v. 44, p. 396–400.

Tilly, H.P., B.J. Gallagher, and T.D. Taylor, 1982, Meth-ods for correcting porosity data in a gypsum-bear-ing carbonate reservoir: Journal of PetroleumTechnology, v. 34, p. 2449–2454.

Von Bergen, D., and A.V. Carozzi, 1990, Experimen-tally-simulated stylolitic porosity in carbonate rocks:Journal of Petroleum Geology, v. 13, p. 179–192.

Wanless, H.R., E.A. Burton, and J.J. Dravis, 1981,Hydrodynamics of carbonate fecal pellets: Journalof Sedimentary Petrology, v. 51, p. 27–36.

Ward, R.F., C.G.St.C. Kendall, and P.M. Harris, 1986,Upper Permian (Guadalupian) facies and theirassociation with hydrocarbons—Permian Basin,West Texas and New Mexico: AAPG Bulletin, v. 70,p. 239–262.

Page 260: Reservoir Quality Prediction in Sand and Carbonates

249

Chapter 16

Depositional Controls Over PorosityDevelopment in Lithic Sandstones

of the Appalachian Basin:Reducing Exploration Risk

Richard SmosnaKathy R. Bruner

Department of Geology and Geography, West Virginia UniversityMorgantown, West Virginia, U.S.A.

ABSTRACT

Litharenites and sublitharenites of the Devonian Lock Haven Formationcontain abundant rock fragments of shale and phyllite. These labile grainssuffered varying degrees of destruction in several depositional environ-ments; hence, sedimentary processes largely controlled the sandstones’ min-eral composition. Fluvial sandstones have a high lithic content, distributarymouth-bar and offshore-shelf sandstones have an intermediate content, andbarrier-island sandstones have a low content.

Primary porosity relates inversely to compaction of the lithic grains,decreasing from a maximum minus-cement porosity of φmc = 33% down tozero as lithics increase. The majority of primary porosity, however, has beenoccluded by cementation. Secondary porosity, created chiefly by dissolutionof the chemically unstable rock fragments, is greatest (φrf = 13%) for sand-stones of a moderate lithic content.

Because of these relationships among depositional processes, lithology,and porosity, we predict that sandstones of different sedimentary environ-ments should exhibit distinct porosity volumes and vary in their reservoirpotential. Mouth-bar sandstones will have good total porosity, good sec-ondary porosity, and offer the best reservoir quality. Shelf sandstones willhave fair total porosity, most of which is secondary, whereas beach sand-stones will have low total porosity, most of which is primary. Fluvialsandstones will be the poorest reservoirs.

Smosna, R., and K.R. Bruner, 1997, Depositional con-trols over porosity development in lithic sandstonesof the Appalachian Basin: reducing exploration risk,in J.A. Kupecz, J. Gluyas, and S. Bloch, eds.,Reservoir quality prediction in sandstones and car-bonates: AAPG Memoir 69, p. 249–265.

Page 261: Reservoir Quality Prediction in Sand and Carbonates

250 Smosna and Bruner

INTRODUCTION

The majority of secondary porosity in sandstones isattributed to the dissolution of feldspars and carbon-ate minerals (Heald and Larese, 1973; Pittman, 1979;Schmidt and McDonald, 1979a, b; Björlykke, 1983,1984; Shanmugam, 1985, 1990). In contrast, substantialsecondary porosity in Upper Devonian reservoirs ofthe Appalachian basin has been generated by theleaching of metamorphic and sedimentary rock frag-ments (Bruner and Smosna, 1994). Moreover, we haveobserved lithmoldic porosity in sandstones through-out the Appalachian stratigraphic section from theLower Devonian to the Pennsylvanian. Other workershave noted comparable dissolution porosity in sand-stones, including major Tertiary reservoirs of the GulfCoast such as the Wilcox Group and Frio Formation(Moncure et al., 1984; Siebert et al., 1984; Parnell, 1987;Stonecipher and May, 1990).

We suspect that lithmoldic porosity may be evenmore widespread, although perhaps overlooked ormisidentified. Sediments derived from collision-suturebelts or foreland fold and thrust belts contain an abun-dance of chemically unstable metamorphic and sedi-mentary rock fragments (Dickinson and Suczek, 1979;Bird and Molenaar, 1992; Dickinson, 1988; Potocki andHutcheon, 1992). Likewise, sands of large river systemsthat drain passive, Atlantic-type continental margins areenriched in metamorphic and sedimentary rock frag-ments (Potter, 1978; Loucks et al., 1984). Sandstones ofthese geologic settings, in particular, may be most sus-ceptible to the creation of similar secondary porosity.

In Upper Devonian reservoirs of the Appalachianbasin, creation of secondary porosity was controlled bythe number of chemically unstable grains, the amountof primary porosity, and the rocks’ early permeability.Primary porosity, in turn, had been influenced by com-paction and the volume of ductile grains, quartz over-growths, dolomite cement, authigenic clays, and theintroduction of solid bitumen (Bruner and Smosna,1994). Figure 1 traces the changing petrographic com-position of these reservoir sandstones through time:from deposition, through compaction, to grain dissolu-tion. Mechanical compaction and cementation reducedprimary porosity to ~1%, whereas chemical leachinglater generated 5% secondary porosity.

Knowledge of the exact relationship between poros-ity and petrographic composition may allow for theprediction of porosity distribution in developmentwells. But such correlations generally offer little aid forexploratory wells. In sparsely tested areas, geologistsmay have few data concerning the content of solubleminerals, ductile minerals, cement volumes, percentclay, or presence of bitumen in a target sandstone.Instead, predicting porosity in advance of the drillmust be linked to the kinds of information available atan early stage of exploration, such as depositional envi-ronment (Pryor, 1973; Loucks et al., 1984; Stonecipheret al., 1984; Stonecipher and May, 1990), hydrogeologi-cal regime (Björlykke, 1983; Galloway, 1984; Shan-mugam, 1990), or burial depth (Sclater and Christie,1980; Chilingarian, 1983; Baldwin and Butler, 1985).

Upper Devonian sandstones of the Catskill deltaiccomplex constitute a major exploration play for nat-ural gas in Pennsylvania. Recent activity has focusedon the Lock Haven Formation of Centre and Clinton

Figure 1. Diagenetic pathway for Lock Haven lithicsandstones showing petrographic composition atthe time of deposition, after compaction, and afterchemical leaching. Black = other minerals, cmt =cement, φp = primary porosity, φs = secondary porosi-ty from the dissolution of rocks fragments andfeldspars, Q = sedimentary quartz, RF = rock frag-ments, F = feldspars.

Page 262: Reservoir Quality Prediction in Sand and Carbonates

counties, ~900 m thick and consisting of interbeddedshale, sandstone, and siltstone (Figure 2). Many of thesandstones, however, have a low porosity and perme-ability; they possess only modest storage capacity andare marginally profitable reservoirs. It is especiallyimportant, therefore, under such economic conditions,that new gas prospects be evaluated and appraisedjudiciously. In this chapter, we document a semiquan-titative relationship between observed porosity andinferred depositional facies, a relationship that allowsfirst-order prediction of primary and secondary porevolumes before drilling.

Lock Haven sandstones of the present study go bythe drillers’ informal member names of Warren, Speech-ley, Bradford, and Elk. Fifty-one samples were investi-gated by thin-section microscopy [five more thandescribed in Bruner and Smosna (1994)]; in addition,several of these were investigated by scanning electronmicroscopy and the X-ray diffraction method. The sam-ples, taken from sidewall and full-bore cores, came fromfive wells in Clinton County, two in Centre County, and

one in Somerset County (Figure 3). The paleoshoreline isthought to have passed through central Pennsylvania atthis time (Dennison, 1985), and these sandstones repre-sent a mix of terrestrial, transitional, and shallow-marineenvironments. Porosities range from 0% to 20%, andhorizontal air permeabilities from <0.001 to 6 md.

SEDIMENTARY ENVIRONMENTS

Three full-bore cores of Lock Haven sandstoneswere recovered in Clinton County and analyzed petro-graphically. Well numbers, exact locations, and mem-ber names, however, are proprietary information andcannot be released. We described these cores in detail,noting lithologies, textures, structures, and verticaltrends in order to identify and interpret the severalfacies. Facies interpretations have been confirmed bycomparison with standard, well-recognized sedimen-tary models (Elliott, 1978a, b; Johnson, 1978; Bouma etal., 1982; Coleman and Prior, 1982; McCubbin, 1982;Miall, 1984; Reinson, 1984; Walker, 1984). The cored

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 251

Figure 2. Upper Devonianstratigraphy of Clinton County,Pennsylvania, includes the900-m-thick Lock HavenFormation. This formation con-sists of several packages ofsandstones (four of which weresampled for this study: Warren,Speechley, Bradford, and Elk)disseminated through a thickinterval of shale and siltstone.

Page 263: Reservoir Quality Prediction in Sand and Carbonates

252 Smosna and Bruner

sandstones compose a systems tract of distributarymouth bars, offshore sand ridges, and barrier islands.A fourth, fluvial facies is present in the systems tract,but only sidewall cores are available. Facies were com-pared to geophysical well logs, particularly thegamma-ray signatures, and our depositional interpre-tations could then be extended throughout the region,based on an integration of core analyses, gamma-raycorrelations, and sand-body geometries.

Distributary Mouth Bar

The mouth-bar sandstone, 8.5 m thick, constitutes acoarsening-upward sequence, and its funnel-shapedgamma-ray signature suggests a correspondingdecrease in clay content (Figure 4). Rocks of the lowerpart consist of alternating very fine sandstone and siltyshale, and the base grades downward into an underly-ing shale. One sandstone at the base contains abundantshale clasts and carbonate-cemented nodules. Towardthe top, the section is made up of fine and very finesandstone with no shale. Wavy to lenticular beddingand current-ripple bedding are the prevalent sedimen-tary structures in the lower part, whereas parallel lami-nae become predominant above. Body fossils (pyritizedbrachiopods) are rare throughout the unit, althoughsand- and clay-filled burrows are common in the shales.Ripple cross-laminae, parallel laminae, shale partings,rip-up clasts, and burrows occur at the very top.

This distributary mouth-bar sandstone accumu-lated during the westward progradation of a LockHaven delta front across the shelf. The lower sectionwas deposited on distal reaches of the mouth bar near

normal wave base. At that depth, deposition frequentlyalternated between mud and sand. Sedimentation rateof the mud was slow, allowing extensive burrowing byinfauna, and shale clasts testify to occasional periodswhen currents eroded the muddy sea floor. The distalmouth-bar sandstone passes upward in the core to aproximal facies. The upper coarser grained sectionformed in a somewhat higher energy setting when thebar built itself up to sea level. An increasing energylevel is also denoted by the vertical trend in sedimen-tary structures. Fine sandstone with parallel laminaereflects deposition within the swash zone along thecrest of the bar. Very fine, rippled, and burrowed sand-stone formed on the back of the mouth bar, wherewave action was slightly reduced from that of the crest.

Beneath the mouth-bar sandstone lies a prodeltaicshale with minor siltstone and very fine sandstone.Shales are silty, organic rich, micaceous, and well lam-inated. The few burrows are mostly small and hori-zontal; they are filled with silt and sand, and containpyrite. Several large sand-filled Rhizocorallium tracesare also identified. The siltstones and sandstones occurin thin beds, isolated laminae, or starved ripples, andthe main structure is small-scale cross-laminae.Thicker beds may have a scour base and be parallellaminated. Deformation structures developed wherethe underlying muds were soupy and weak: loadstructures along the base of sandstones, distortion ofbedding around large burrows and sandstone lenses,and slump structures with small-scale folding andbrecciation. Some sandstones contain vertical escapeburrows; others are extensively bioturbated.

Figure 3. Generalized paleo-geography of Pennsylvania atthe end of the Frasnian (afterDennison, 1985). The shore-line passed through or nearCentre and Clinton counties,separating eastern rivers anddeltas of the Catskill complexfrom a shallow western sea.Cores of this study were takenfrom eight wells in Clinton,Centre, and Somerset coun-ties. The inset map shows thelocation of Pennsylvania(black) and the Appalachianbasin (stippled) in the easternUnited States.

Page 264: Reservoir Quality Prediction in Sand and Carbonates

Above the mouth-bar sandstone is a black shale andshaly siltstone. These rocks resemble the shales andsiltstones beneath, and they attest to a rise of sea levelor switching of distributary channel and a return ofthe prodeltaic environment.

Offshore Sand Ridge

Although subtle, the sand-ridge facies (8.7 m thick)displays a coarsening-upward texture and a funnel-shaped gamma-ray signature (Figure 5). The unit’s base

consists of siltstone with minor very fine sandstone andshale. Parallel lamination is the dominant structure,although ripple trough cross-laminae are also present.Upward in the core, very fine sandstone and siltstonewith parallel laminae, planar cross-laminae, and rippletrough cross-laminae make up the middle section.Scour surfaces and shale clasts are present, as are anumber of storm deposits, noted by their sharp basalcontact, cross-laminae, and burrowed cap (Kreisa,1981). Very fine sandstones with parallel laminae mark

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 253

Figure 4. Columnar section of distributary mouth-bar sandstone illustrates the associationof lithologies, textures, structures, and vertical facies associations. The gamma-ray log has afunnel shape.

Page 265: Reservoir Quality Prediction in Sand and Carbonates

254 Smosna and Bruner

the top of the unit. Zones of reworked fossils and shaleclasts occur there. Skeletal grains, including bra-chiopods, crinoids, and gastropods, are common in thesandstones and siltstones from base to top, but burrowsare confined to shale partings.

These fossiliferous sandstones formed as a sandridge on the offshore Lock Haven shelf. As the ridgeaggraded into shallower water, wave and currentactivity increased, and grain size coarsened slightly.Simultaneously, the predominant sedimentary struc-ture changed from settle-out lamination (no flow) tocross-lamination (low-flow regime) to parallel lamina-tion (upper-flow regime). Occasional periods of sea-floor erosion are indicated by the scour surfaces,rip-up clasts, thin tempestites, and fossil lags. Numer-ous epifaunal invertebrates inhabited the sand ridgeand surrounding shelf; however, infaunal organismsburrowed into the sediment only during breaks insand deposition (shale partings).

Underlying the sand-ridge sandstone is a burrowedsiltstone and shale with minor sandstone, interpretedto have accumulated on the deep shelf below wavebase. Near-absence of body fossils in the siltstones

with only moderate burrowing points to a stressedbiological environment; the water chemistry was per-haps dysaerobic. Organic matter, pyritized fossildebris, and the green-gray color provide additionalevidence for reducing conditions at the silt–waterinterface. Interbedded shales are well laminated andhave few burrows, suggesting an even lower oxygenlevel in somewhat deeper water. Very fine sandstones,on the other hand, are fossiliferous (crinoids, bra-chiopods, bivalves, and tentaculitids) and moderatelyburrowed, reflecting deposition near wave base andunder oxygenated conditions. Above the sand-ridgesandstone, the overlying facies suggests a drowning ofthe shelf by transgression and a return to the deposi-tion of similar deep-water fine clastics.

Barrier Island

We divide the 11.6-m stratigraphic section intothree parts (Figure 6). The middle part contains a dis-tinctive parallel- and cross-laminated, very coarsesandstone. Particles include coarse rock fragments,feldspars and quartz, large skeletal debris, and phos-phate grains. The basal contact is sharp. Overlying this

Figure 5. Columnar section ofoffshore sand-ridge sand-stone. The sandstone is fossil-iferous and parallellaminated; overlying andunderlying siltstone andshale are bioturbated.Gamma-ray log shows a subtle funnel shape. Thesymbols are the same as inFigure 4.

Page 266: Reservoir Quality Prediction in Sand and Carbonates

is a minor shaly, sandy siltstone and shale that is bio-turbated and displays some deformation structures.Stratigraphically higher lies a fossiliferous sandstone:very fine to fine grained with abundant brachiopods,gastropods, crinoids, bivalves, tentaculitids, and bry-ozoans. Predominant structures are ripple bedding,burrows (especially Asterosoma), and shale partings.

The upper section of the core consists of interbeddedsiltstone, very fine sandstone, and shale. Sandstonesand siltstones are somewhat fossiliferous, containingbrachiopods, crinoids, and ostracods, and the primarystructures include both parallel and cross-lamination.Also observed are a small-scale scour surface and a sin-gle thin bed of brachiopod valves with concave-downorientation. The shales may be laminated or burrowed.Some beds have been contorted by soft-sediment defor-mation, and many possess abundant organic matter.

The lower part of the core is marked by an exten-sively bioturbated fine sandstone. Coarse quartz sandis scattered throughout, and hints of laminae appearwhere burrowing is less intense. Very fine sandstone,siltstone, and shale alternate at the very bottom. Thesandstones and siltstones occur in thin beds, laminae,and lenses, which show ripple bedding and contain afew small brachiopods. In contrast, the shales are wellburrowed, and burrows illustrate a diverse assemblageof forms: sand- and silt-filled, small and horizontal,long and vertical, vertical U-shaped, and horizontalAsterosoma.

This sandstone body represents a transgressive bar-rier-island complex. We interpret an erosional ravine-ment to be present at the base of the middle section,having formed during a time of shoreface retreat (Swift,1975). The barrier facies (that is, upper-shoreface andforeshore sands) were removed by erosion when sealevel rose, and waves of the surf zone attacked thebeach. Very coarse sandstone accumulated as a lagdeposit atop the ravinement surface. Above the lag, rip-pled, fossiliferous finer grained sandstone with shaleformed in the middle-lower shoreface environment asthe transgression continued. The overlying, still-finersiliciclastics were laid down on the inner shelf, markinga transition from nearshore sand to offshore mud. Waveand current energy decreased systematically offshore,and grain size of the sediment concurrently decreased.This fining-upward sequence of transgressivebeach–inner shelf sandstones is paralleled by the bell-shaped gamma-ray signature (Figure 6). Storms infre-quently reworked the shelf sediment, as evidenced bythe scour structure and resedimented brachiopods.

Lagoonal sediments lie beneath the barrier-islandcomplex. Laminated, fine sandstone with scatteredcoarse quartz is thought to have been deposited by theaction of overwash onto the leeward side of the barrierisland. Storms breaking over the barrier transportedsand into the lagoon; extensive bioturbation then devel-oped during the interval following rapid deposition.Sediments of the lagoon proper included sparsely fossil-iferous, very fine sand and thoroughly burrowed mud.

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 255

Figure 6. Columnar section of a transgressive barrier-island complex is marked by the coarse lag deposit ofshoreface retreat. Underlying lagoonal sandstones are extensively bioturbated, whereas overlying sandstonesof the lower shoreface and inner shelf are laminated and fossiliferous. Gamma-ray log has a relatively sharpbase and bell shape. The symbols are the same as in Figure 4.

Page 267: Reservoir Quality Prediction in Sand and Carbonates

256 Smosna and Bruner

Streams

Fluvial sandstones are interpreted by their blockyor spiky pattern on the gamma-ray log, the shoestringgeometry of sandstone on isopach maps, and an orien-tation perpendicular to the paleoshoreline (Boswelland Jewell, 1988). Sidewall cores show the rocks to betexturally mature and unfossiliferous.

FACIES AND LITHOLOGIES

Two thirds of the 51 sandstones are classified assublitharenite, and the remainder as litharenite(Bruner and Smosna, 1994). Although the litharenitescontain a greater number of rock fragments, they areotherwise quite similar to the sublitharenites. The twogroups in fact form a petrographic continuum of lithic-rich rocks with a mean quartz-feldspar-rock fragment(Q/F/RF) ratio of 81/2/17 (Figure 7). Sorting is gener-ally moderate to good. Mean grain size varies from0.100 mm for shelf sandstones to 0.130 mm for mouth-bar sandstones, to 0.170 mm for fluvial sandstones,and to 0.240 mm for beach sandstones.

Quartz includes normal, undulose, polycrystalline,and stretched varieties, and many grains have inclu-sions of vermicular chlorite, kaolinite, and zircon, orintergrowths of feldspar. Rock fragments are over-whelmingly a mixture of phyllite and shale (Figure 8A).Phyllite clasts are composed of fine mica and chloritewith minor amounts of quartz, feldspar, apatite, andgarnet; they display a marked foliation. Shale clasts con-sist of illite, quartz silt, and kaolinite. Rock fragments ofdolomite, sandstone, and volcanics are, by contrast,very rare. Most of the feldspars appear fresh, but someare partly altered to sericite and illite, or replaced bydolomite. Muscovite, biotite, and chlorite constitute themain accessory minerals. All but six samples have aclay content <10% of the total rock volume.

During diagenesis, many rock fragments andfeldspars were removed by dissolution, but secondarymoldic pores can readily be differentiated in thin sec-tion, interpreted as to their origin, and point-counted.Rock fragments have also been squeezed into apseudomatrix, but pseudomatrix can be identified andcounted. These values for leached rock fragments,leached feldspars, and compacted rock fragments arethen added to data for existing grain percents to deter-mine the mineral composition before dissolution; thatis, the original content of rock fragments and feldspars.Recalculating point-count data for all samples yields anoriginal Qo/Fo/RFo of 77/3/20 (Table 1).

The source area for these Upper Devonian sand-stones must have contained a mixture of metamorphicand sedimentary rocks. Lithic-rich lithologies of theLock Haven Formation match the composition pre-dicted by Dickinson and Suczek (1979) and Dickinson(1988) for recycled sands from an orogenic prove-nance. This mix of sediments came from uplifted strataof the Acadian Mountains to the east, particularlyLower Cambrian low-grade quartzites, phyllites, andslates and their sedimentary cover. Continent-continent collision between North America, westernEurope, and the Avalone terrane had just recently cre-ated the Acadian fold and thrust belt, and the Catskilldeltaic complex constitutes the siliciclastic wedge thatfilled the adjoining foreland basin (Ettensohn, 1985).

The 51 samples obviously form a single populationof sandstones, but the four sedimentary facies illustratesomewhat different mineral compositions (Figure 9).Despite an overlap, they have different contents of rockfragments (RFo before leaching). Mineral compositionwas mainly controlled by mechanical destruction ofrock fragments in the depositional environment (Daviesand Ethridge, 1975; Mark, 1978; Espejo and López-Gamundí, 1994). Breakdown of shale and phyllitegrains produced fine detritus of micas and clays; hence,muscovite, biotite, chlorite, illite, sericite, and smectiteare common components of both Lock Haven rock frag-ments and the sandstones’ matrix. Other controls onsandstone mineralogy, such as tectonism, provenance,and climate, presumably remained constant in centralPennsylvania during Lock Haven deposition.

Meandering streams draining the Acadian Moun-tains transported and deposited sands rich in rockfragments (Figure 9). Resultant sandstones have ahigh, although variable, lithic content: RFo rangesbetween 6% and 42% of the total rock volume andaverages 24%. In between terms of compositionalmaturity, these are the most immature samples. Lithiccomposition not only varies considerably across thisfacies, as indicated by a large standard deviation, but therange in lithic content is also significant within any sin-gle fluvial sandstone. For example, the lithic content inone sandstone body varies between 11% and 42% within1 m of section; another, between 6% and 32% in 3 m ofcore; and yet another, between 11% and 36% in 16 m.This extreme mineral variation most likely reflects dras-tic fluctuations in stream velocity and abrasive capacity;furthermore, no other facies exhibits such drasticmineral changes.

Figure 7. Quartz-feldspar-rock fragments ternarydiagram plotting 51 Lock Haven sublitharenites(5% ≤ RF < 25%) and litharenites (RF ≥ 25%).

Page 268: Reservoir Quality Prediction in Sand and Carbonates

Abundant rock fragments were then delivered bystreams to distributary mouth bars at Lock Haven deltafronts. Mouth-bar sandstones have a notably less vari-able lithic content, and RFo of these sandstones ranges

between 7% and 24%; the mean value is 15%. Presum-ably, these sediments were subjected to intense waveaction and grain disintegration on the delta front. Fromthe delta, lithic sands were swept seaward by storm cur-rents to offshore sand ridges. Shelf sandstones have amean RFo close to that of mouth-bar sandstones, 14%;the range of lithics is somewhat lower (4%–21%). Stormaction and offshore swells thus produced only minordestruction of labile rock fragments on the sand ridges.Barrier-island sands, worked continuously by longshorecurrents and breaking waves, possess the highest com-positional maturity. Rock fragments are reduced to amean value of 7%, and the narrow range of RFo(4%–12%) indicates widespread breakdown of labilegrains in this high-energy depositional setting. Differ-ences in lithic content between the four facies do exist,depending to a large degree on the relative survivabilityof these unstable grains.

Results of this study compare favorably with otherpublished investigations: mineral composition of silici-clastic sediments is sensitive to the depositional environ-ment (Cameron and Blatt, 1971; Stonecipher and May,1990). Fluvial sandstones are richest in metamorphicand sedimentary rock fragments; mouth-bar sandstonescontain, on average, less than two thirds of the lithic con-tent of fluvial facies; rock fragments are reduced but justa few percentage points between mouth-bar and shelfsandstones; the low content of rock fragments in barrier-island sandstones is half that of the mouth-bar facies. Intotal, two thirds of the rock fragments were destroyedbetween Lock Haven rivers and barrier islands. In likemanner, Harper and Laughrey (1987), studying slightlyyounger Devonian Venago sandstones in Pennsylvania,reported a 20% decrease in rock fragments (pluspseudomatrix) between fluvial and mouth-bar sand-stones, and a virtual elimination of rock fragments inforeshore sandstones. Davies and Ethridge (1975)observed a three-quarter reduction between fluvial anddeltaic sandstones of the Eocene Wilcox Group in Texas.

FACIES AND PRIMARY POROSITY

Approximately one third of the porosity in the 51sandstones is primary (Figure 8B), ranging up to a maxi-mum value of 13%. Primary pores between rigid quartzgrains are triangular in cross section, although whenadjacent to ductile rock fragments, they can be quiteirregular. Their size as measured in thin section variesbetween 0.005 and 0.080 mm, typically about one fourthto one third of the diameter of surrounding grains.

Because sedimentary environment controlled thecontent of rock fragments to a large extent, and becausesandstone lithology in turn controlled the degree ofcompaction, samples of this study reveal a close associ-ation between facies, mineral content, compaction, andprimary porosity (Figures 10, 11). Compactiondeformed the ductile rock fragments of shale andphyllite and squeezed them into much of the inter-granular pore space (Figure 8A). As a result, porositysystematically decreases in sandstones with anincreasingly higher lithic content (Rittenhouse, 1971;Smosna, 1989). The graph in Figure 10 illustrates this

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 257

Figure 8. Photomicrographs of Lock Haven sand-stones. Bar scales equal 0.250 mm. (A) Sandstonewith abundant rock fragments of phyllite and shale(brown grains). Compaction has deformed these duc-tile lithics, squeezing them into intergranular porespace. (B) Large and small triangular primary poresbetween quartz grains have been lined and partlyoccluded by solid bitumen (black). (C) Oversizedmoldic pores >0.100 mm (center) were generated bythe dissolution of rock fragments.

Page 269: Reservoir Quality Prediction in Sand and Carbonates

258 Smosna and Bruner

Table 1. Petrographic Data, Lock Haven Sandstones.*

SampleNumber Facies Qo Fo RFo φmc φp φrf

1 FL 40 1 21 0 0 02 FL 74 5 16 5 1 63 FL 69 7 15 2 0 84 FL 59 3 29 5 0 25 FL 63 7 6 23 3 06 FL 50 1 32 13 0 07 FL 69 4 11 8 2 48 FL 51 3 42 0 0 09 FL 65 5 11 13 2 6

10 FL 53 2 38 5 0 011 FL 46 3 36 9 0 012 FL 51 5 31 8 0 013 FL 57 3 32 3 0 014 FL 57 2 36 1 0 215 FL 59 5 28 2 0 116 FL 58 2 12 27 0 017 FL 67 2 15 15 0 018 FL 59 1 26 10 0 019 FL 52 2 27 16 0 020 MB 62 4 20 12 0 021 MB 70 3 18 8 2 1322 MB 61 13 15 10 2 1123 MB 54 2 24 4 0 124 MB 44 1 7 0 0 025 MB 68 7 12 13 6 226 MB 67 2 14 12 0 627 MB 66 1 11 11 3 528 MB 69 1 8 20 0 029 MB 57 0 24 8 0 030 MB 65 2 10 19 0 031 SR 63 9 21 6 2 432 SR 49 12 21 10 5 833 SR 60 3 11 17 0 034 SR 58 1 21 5 0 635 SR 51 1 21 10 0 136 SR 67 3 9 19 2 137 SR 67 2 5 25 0 038 SR 73 1 4 20 0 039 SR 35 1 12 5 0 040 B 34 2 4 33 0 041 B 71 0 6 20 0 042 B 75 0 9 12 3 343 B 65 3 12 13 0 244 B 71 1 8 15 0 045 B 81 0 5 11 1 146 B 74 2 4 12 0 047 B 69 1 5 17 7 348 B 72 0 7 24 3 249 B 59 1 7 32 13 150 B 56 1 10 32 1 651 B 62 1 8 11 0 0

*B = beach/barrier island, FL = fluvial, MB = distributary mouth bar, SR = offshore sand ridge, F = feldspar, Q = quartz, RF = rock fragments.

Page 270: Reservoir Quality Prediction in Sand and Carbonates

correlation between lithic content and minus-cementporosity (sum of remaining primary porosity, quartzand dolomite cements, plus bitumen). Minus-cementporosity equals the volume of intergranular porespace before cementation but after the completion ofmechanical compaction. Maximum minus-cementporosity of φmc = 33% occurs in a sublitharenite with anoriginal lithic content RFo of 4%, and porosity falls tozero in a litharenite with a lithic content of 42%. (Twoother sandstones with excessive clay matrix also pos-sess a φmc = 0.) The function φmc = 150/RFo approximatesthe correlation between mineral content and originalintergranular porosity.

In all four facies, most available pore space aftercompaction has been occluded by cement, matrix, andauthigenic minerals (Bruner and Smosna, 1994).Although minus-cement porosity reaches 33%, finalprimary porosity φp does not exceed 13%, and twothirds of the samples have none. Syntaxial quartzcement has overgrown many detrital quartz grains.Ferroan dolomite commonly replaced feldspars androck fragments, and in many places it extends beyondthe boundary of framework grains to occupy adjacentpore space as a void-filling cement. Authigenic clays(illite, chlorite, and kaolinite) precipitated within inter-granular pores and pore throats. Some clay matrix alsooccupies pore space. Finally, solid bitumen occurs asglobs and interstitial stringers that coat detrital grainsand line pores. Figure 10 also illustrates the present,reduced primary porosity. All but three samples plotbelow the curve φp = φmc/3; that is, more than two thirdsof the original porosity has been occluded.

Minus-cement porosity can be related to sedimen-tary facies and mineral composition (Figure 11). Fromthe fluvial environment to distributary mouth bar tosand ridge to barrier island, the lithic content of sand-stones was continuously reduced as mechanicalprocesses destroyed more and more shale and phylliteclasts. Fewer ductile rock fragments resulted in a lowerdegree of compaction. As a consequence, the averageminus-cement porosity rises from 9% in fluvial sand-stones to 11% in mouth-bar sandstones, to 13% in sand-ridge sandstones, and finally to 19% in barrier-islandsandstones. Reduced porosity after cementation, how-ever, is a different story, being very low in all but oneLock Haven facies. Present porosity averages 0.4%,1.2%, and 1.0% in fluvial, mouth-bar, and shelf sand-stones, respectively. Half of the barrier-island samples,on the other hand, have some measurable thin-sectionporosity; mean porosity for these 12 samples is 2.3%(two to six times more than in the other facies); and themaximum reaches 13%.

FACIES AND SECONDARY POROSITY

Two thirds of the porosity in the sandstones is sec-ondary, and most of this is attributed to the leaching ofphyllite and shale rock fragments (Figure 8C). Com-plete grain dissolution produced oversized molds of0.100 to 1.700 mm. These molds often mimic compactedrock fragments in both size and shape, indicating thatchemical leaching occurred after compaction. Intra-granular pores of 0.025 to 0.050 mm resulted from par-tial leaching of rock fragments, whereas intragranular

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 259

Figure 9. Graph of RFo (content of rock fragments before chemical leaching) sedimentary environment; RFoequals the volume percent of existing lithic grains plus pseudomatrix plus secondary porosity interpreted tohave formed by dissolution of lithics. Vertical bars represent mean value for each facies. The data exemplify afacies dependency of lithic content among Lock Haven sandstones: lithic grains progressively decreased alongthe depositional systems tract.

Page 271: Reservoir Quality Prediction in Sand and Carbonates

260 Smosna and Bruner

micropores (<0.005 mm) illustrate a beginning stage ofgrain leaching. Partial and complete dissolution offeldspars has created additional secondary porosity.

The relationship between rock composition and sec-ondary porosity is illustrated in Figure 12 (bottom partof graph), in which the upper limit of lithmoldicporosity defines a bell-shaped curve. Samples with alow content of rock fragments would, of course, con-tain few chemically unstable grains to remove byleaching. Two lithic-poor sandstones, for example,with original contents of 5% and 9% rock fragments,have 3% lithmoldic porosity. At a maximum, then,about half of the shale and phyllite rock fragments inthese samples have dissolved. On the other hand,litharenites with >21% rock fragments never devel-oped significant lithmoldic porosity. Compaction wasextreme for sandstones with a high volume of lithics:mean values for primary porosity and permeabilitydropped to 0.4% and 0.04 md, and leaching fluidscould not enter the rock to dissolve the unstablegrains. Accordingly, a very small number of lithicgrains were leached, and secondary porosity inlithic-rich sandstones does not exceed 2%.

Secondary porosity is greatest (φrf = 13%) in sand-stones with an intermediate content of unstable grains.With an intermediate composition, compaction wasmoderate, and some primary porosity remained in thedeep subsurface. Primary porosity not only con-tributed to reservoir quality, but allowed the introduc-tion of leaching fluids, which dissolved rockfragments to create the secondary porosity. In theseintermediate sandstones, as many as three quarters ofthe lithic grains have been dissolved.

Figure 12 also underscores the relationship betweensecondary porosity and sedimentary facies. Secondary

porosity is greatest (φrf values of 6%–13%) in those sub-litharenites with an original content of rock fragmentsbetween 10% and 21%. The four facies, however, havedifferent mineral compositions, leading to differencesin the distribution of lithmoldic porosity. Two thirdsof the mouth-bar sandstones and two thirds of theshelf sandstones initially contained an intermediateamount of unstable grains (10% ≤ RFo ≤ 21%). In con-trast, less than one third of the fluvial sandstones hadan intermediate volume of unstable grains; this faciesincludes many samples with an overabundance oflithics. And very few of the barrier-island sandstoneshad an intermediate content; most contained too fewlithics. Again, composition of the sandstones waschiefly a function of mechanical destruction of rockfragments in the depositional environment. Corre-spondingly, the mean value of secondary porosity is3.5% in the mouth-bar sandstones, 2.2% in the sand-ridge samples, and 1.5% in both the fluvial and bar-rier-island sandstones.

In a recent analysis of Carboniferous reservoir rocksin Australia, Hamlin et al. (1996) reached comparableconclusions. They identified four sandstone facies in anonmarine braid-delta system; the rocks are classifiedas sublitharenites, lithic grains range in abundancefrom 2% to 40%, and most of these are clasts of meta-morphic and sedimentary rocks. Porosity is chieflysecondary, originating from the dissolution of rockfragments and rare feldspar. Their data show that asthe lithic content increases from 13% to 21%, sec-ondary porosity actually falls from 3.8% to 2.4%. Max-imum porosity occurs in sandstone of an intermediatelithic composition.

Stonecipher and May (1990) also identified an anal-ogous relationship between depositional facies and

Figure 10. Graph of RFominus-cement porosity andreduced primary porosity. φmc= existing primary porosityplus quartz and dolomitecements plus bitumen, andrepresents intergranularporosity before cementation.Porosity loss by compactionas a function of increasinglithic content is approximat-ed by φmc = 150/RFo. φp =reduced porosity aftercementation, and in mostsamples more than twothirds of the minus-cementporosity has been occluded.

Page 272: Reservoir Quality Prediction in Sand and Carbonates

secondary porosity in the Wilcox Group of Texas.Although most porosity is primary in origin, sec-ondary porosity in these sublitharenites (lithics rangefrom 8% to 51%) depends in part on the content of rockfragments. Secondary porosity reaches highest values(3%–6%) in sandstones of distributary channel andmouth-bar facies with an intermediate content of rockfragments (17%–22%). Lithic-poor sandstones of theshoreface and tidal-flat facies, as well as lithic-richsandstones of the tidal-channel facies, possess little orno secondary porosity.

Core Laboratories measured horizontal air perme-abilities for 34 Lock Haven samples of this study, and27 have a value <0.10 md; these are truly tight sand-stones (unpublished report to Eastern States Explo-ration Company). But of the remaining samples—withpoor to fair permeabilities (0.10–5.85 md)—fluid flowmay be linked to porosity type and amount. Networksof lithmoldic pores are commonly observed in thin sec-tions where several neighboring rock fragments havebeen leached. These networks, aligned parallel to bed-ding, must surely enhance the rocks’ horizontal perme-ability. Also, the small intergranular primary poresmay act as conduits between larger lithmoldic pores.The best reservoir sandstones, therefore, seem to pos-sess a high total porosity (primary plus secondary) andan intermediate to high secondary porosity.

POROSITY PREDICTION

Subsurface exploration for Upper Devonian reser-voirs in the Appalachian basin is based primarily on themapping of sandstone members as identified by geo-physical well logs. Prospects are then definedby the geometry of a sandstone body, its thickness, ori-entation, association with other sandstones, positionwith respect to shoreline or shelf margin, and structuralconfiguration. Figure 13 depicts a composite isopachmap of genetically related prospect sandstones: all four

facies of this study have been arranged into an idealizedUpper Devonian systems tract (modified from Boswelland Jewell, 1988).

The basic premise for our porosity predictions isthat measurable differences in lithic content existamong the four sandstone facies. Each facies shouldconsequently register different values for both pri-mary and secondary porosities. Furthermore, within asandstone member, the porosity values should changeregularly along trends perpendicular to and parallel tothe paleoslope.

The amount of rock fragments will decrease system-atically along the systems tract of transitional-marinesandstones. Fluvial sandstones are compositionallyimmature litharenite with a high volume of lithicgrains. Lithic content, however, can be extremely vari-able, and sublitharenite with comparatively few rockfragments may be closely interbedded with lithic-richsandstone. Compared to this abundance of rock frag-ments in the fluvial facies, lithic grains are reduced byone third on the distributary mouth bar of the deltafront, slightly more on the offshore shelf, and by twothirds on the barrier island.

Compaction of ductile metamorphic and sedimen-tary rock fragments in the litharenites and sublitharen-ites of a foreland basin will significantly reduceintergranular porosity. Minus-cement porosity may beapproximated by the equation φmc = 150/RFo. Usingmean values for Lock Haven samples, primary poros-ity before cementation of fluvial sandstones may havebeen 6%, of mouth-bar sandstones 10%, of shelf sand-stones 11%, and of beach sandstones 21%. These valuesreflect a doubling and redoubling of minus-cementporosity along the systems tract.

Primary porosity, however, is almost everywheregreatly occluded by the introduction of cements, authi-genic clays, and solid bitumen. In fact, almost ninetenths of the minus-cement porosity will eventually bedestroyed. Again, using mean Lock Haven values,

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 261

Figure 11. Graph of minus-cement porosity vs.sedimentary environment.Sandstones with a smallervolume of ductile rock frag-ments underwent less com-paction; consequently, theyretain a higher primaryporosity. Vertical bars repre-sent mean value for eachfacies.

Page 273: Reservoir Quality Prediction in Sand and Carbonates

262 Smosna and Bruner

reduced primary porosity will be 0.6% for fluvial sand-stones, 1.1% for mouth-bar and shelf sandstones, and2.1% for beach sandstones. Thus, only compositionallymature beach sandstones should possess any notableintergranular pore space.

Based on our analysis, secondary lithmoldicporosity should develop best in sandstones with alithic content of 10% ≤ RFo ≤ 21%. Most of the mouth-bar sandstones have such an intermediate mineralcontent, with a mean RFo of 15%. This compositioncorrelates to a maximum lithmoldic porosity of 11%on the bell-shaped plot of Figure 12. We expect,therefore, the mean value of lithmoldic porosity forthis facies to be near 3.7% (a third of the maximum).Most of the offshore-shelf sandstones also have anintermediate composition with a mean RFo of 14%,which equates to a maximum lithmoldic porosity of10% and points to a mean of 3.3% (a third of the max-imum). In like manner, we predict that fluvial sand-stones (RFo = 24%) will have a mean lithmoldicporosity of 1.5%, and beach sandstones (RFo = 7%) amean lithmoldic porosity of 1.2%. Secondary porositywill be low in lithic-rich rocks near the source area,but should more than double along the systems tractto the delta front, where the volume of rock fragmentsbecomes less abundant, and then decrease by twothirds as abrasion and mechanical breakdown removestill more lithics from the sediment.

CONCLUSIONS

To conclude, we rank the different sandstone facieswith respect to reservoir quality. (1) Mouth-bar sand-stones will have the best reservoir potential. Totalporosity should be relatively good (mean of 4.8%) andlithmoldic porosity good, although primary porositywill be low. This combination of total and secondaryporosity may also lead to a fair permeability. More-over, half of the stratigraphic section is expected to beof reservoir quality; that is, having a thin-sectionporosity of ≥6% (half of the mouth-bar samples in ourstudy have this much porosity). (2) Shelf sandstoneswill be similar to, but of a slightly lower quality than,mouth-bar sandstones. Total porosity should be fair(4.4%), and most of this is secondary. Only one third ofthe section may be of reservoir quality. (3) Beach sand-stones will be of an even lower quality. Total porosityshould be poor (3.2%), and most of this is primary.Only one third of the section may be of reservoir qual-ity. (4) Fluvial sandstones will be the poorest reser-voirs. Total porosity (2.1%), lithmoldic porosity, andprimary porosity should all be low. Perhaps just onefifth of the section may be of reservoir quality.

Although differences in total porosities appear to besmall, they become considerable in assessing margin-ally profitable reservoirs. In the comparative evalua-tion of drilling prospects, for example, an 8.0-mmouth-bar sandstone with mean porosity of 4.8% hasthe same gas-storage capacity (porosity thickness orφ×h) as an 8.7-m shelf sandstone with 4.4% porosity, a12.0-m barrier-island sandstone with 3.2% porosity,and an 18.2-m fluvial sandstone with 2.1% porosity.Or from a different viewpoint, a mouth-bar sandstone(half of which is of reservoir quality) has 50% moregas-storage capacity than a shelf or beach sandstone ofthe same thickness (a third of which has reservoirquality), and 150% more storage capacity than a fluvialsandstone (a fifth of which is of reservoir quality). Ourpredictions should, of course, offer immediate benefitto the planning of drilling programs in theAppalachian basin, but we expect that the generalizedrelationships among porosity, lithology, and faciesoutlined by our study may be equally applicable toreservoir lithic sandstones of other basins.

ACKNOWLEDGMENTS

We thank Eastern States Exploration Company,especially Mike Canich and John Humphrey, andCNG Development Company for providing core sam-ples and giving permission to publish the data. NevilleJones, Jon Gluyas, and an anonymous personreviewed the paper and made suggestions for itsimprovement. Alison Hanham and Debbie Bensondrafted the illustrations.

REFERENCES CITED

Baldwin, B., and C.O. Butler, 1985, Compactioncurves: AAPG Bulletin, v. 69, p. 622–626.

Bird, K.J., and C.M. Molenaar, 1992, The North Slope

Figure 12. Graphs of sedimentary environment vs.content of rock fragments before leaching (RFo) andlithmoldic secondary porosity (φrf). Secondary poros-ity is greatest for an original lithic content of 10% ≤RFo ≤ 21% (arrow at top).

Page 274: Reservoir Quality Prediction in Sand and Carbonates

foreland basin, Alaska, in R.W. Macqueen and D.A.Leckie, eds., Foreland basins and fold belts: AAPGMemoir 55, p. 363–393.

Björlykke, K., 1983, Diagenetic reactions in sandstones,in A. Parker and B.W. Sellwood, eds., Sediment dia-genesis: Dordrecht, Reidel Publishing Co., p. 169–213.

Björlykke, K., 1984, Formation of secondary porosity:how important is it? in D.A. McDonald and R.C.Surdam, eds., Clastic diagenesis: AAPG Memoir 37,p. 277–286.

Boswell, R.M., and G.A. Jewell, 1988, Atlas of UpperDevonian/Lower Mississippian sandstones in thesubsurface of West Virginia: West Virginia Geologicaland Economic Survey Circular C–43, 143 p.

Bouma, A.H., H.L. Berryhill, H.J. Knebel, and R.L. Bren-ner, 1982, Continental shelf, in P.A. Scholle and D.Spearing, eds., Sandstone depositional environments:AAPG Memoir 31, p. 281–327.

Bruner, K.R., and R. Smosna, 1994, Porosity develop-ment in Devonian lithic sandstones of theAppalachian foreland basin: Northeastern Geology,v. 16, p. 202–214.

Cameron, K.L., and H. Blatt, 1971, Durabilities of sandsize schist and “volcanic” rock fragments during flu-vial transport, Elk Creek, Black Hills, South Dakota:Journal of Sedimentary Petrology, v. 41, p. 565–576.

Chilingarian, G.V., 1983, Compactional diagenesis, inA. Parker and B.W. Sellwood, eds., Sediment diage-

nesis: Dordrecht, Reidel Publishing Co., p. 57–168.Coleman, J.M., and D.B. Prior, 1982, Deltaic environ-

ments, in P.A. Scholle and D. Spearing, eds., Sand-stone depositional environments: AAPG Memoir 31,p. 139–178.

Davies, D.K., and F.G. Ethridge, 1975, Sandstone compo-sition and depositional environment: AAPG Bulletin,v. 59, p. 239–264.

Dennison, J.M., 1985, Catskill delta shallow marinestrata, in D.L. Woodrow and W.D. Sevon, eds., TheCatskill delta: Geological Society of America SpecialPaper 201, p. 91–106.

Dickinson, W.R., 1988, Provenance and sediment disper-sal in relation to paleotectonics and paleogeographyof sedimentary basins, in K.L. Kleinsphen and C. Paola, eds., New perspectives in basin analysis:New York, Springer-Verlag, p. 3–25.

Dickinson, W.R., and C.A. Suzcek, 1979, Plate tectonicsand sandstone composition: AAPG Bulletin, v. 63,p. 2164–2182.

Elliott, T., 1978a, Deltas, in H.G. Reading, ed., Sedimen-tary environments and facies: New York, Elsevier, p. 97–142.

Elliott, T., 1978b, Clastic shorelines, in H.G. Reading,ed., Sedimentary environments and facies: NewYork, Elsevier, p. 143–177.

Espejo, I.S., and O.R. López-Gamundí, 1994, Sourcevs. depositional controls on sandstone composition

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 263

Figure 13. Composite isopachmap of prospect sandstones inthe Upper Devonian area of theAppalachian basin (modifiedfrom Boswell and Jewell, 1988).Fluvial, distributary mouth-bar, offshore sand-ridge, andbarrier-island facies constitutea genetically related systemstract; each exhibits characteris-tic primary and secondaryporosities. The contour intervalequals 3 m.

Page 275: Reservoir Quality Prediction in Sand and Carbonates

264 Smosna and Bruner

in a foreland basin: the El Imperial Formation (MidCarboniferous-Lower Permian), San Rafael basin,western Argentina: Journal of SedimentaryResearch, v. A64, p. 8–16.

Ettensohn, F.R., 1985, The Catskill delta complex andthe Acadian orogeny: a model, in D.L. Woodrowand W.D. Sevon, eds., The Catskill delta: GeologicalSociety of America Special Paper 201, p. 39–49.

Galloway, W.E., 1984, Hydrogeologic regimes of sand-stone diagenesis, in D.A. McDonald and R.C. Surdam,eds., Clastic diagenesis: AAPG Memoir 37, p. 3–13.

Hamlin, H.S., S.P. Dutton, R.J. Seggie, and N. Tyler,1996, Depositional controls on reservoir propertiesin a braid-delta sandstone, Tirrawarra oil field,South Australia: AAPG Bulletin, v. 80, p. 139–156.

Harper, J.A., and C.D. Laughrey, 1987, Geology of theoil and gas fields of southwestern Pennsylvania:Pennsylvania Topographic and Geologic Survey,Mineral Resources Report 87, 166 p.

Heald, M.T., and R.E. Larese, 1973, The significance ofthe solution of feldspar in porosity development:Journal of Sedimentary Petrology, v. 43, p. 458–460.

Johnson, H.D., 1978, Shallow siliciclastic seas, in H.G.Reading, ed., Sedimentary environments and facies:New York, Elsevier, p. 207–258.

Kreisa, R.D., 1981, Storm-generated structures in sub-tidal marine facies with examples from the Middleand Upper Ordovician of southwestern Virginia:Journal of Sedimentary Petrology, v. 51, p. 823–848.

Loucks, R.G., M.M. Dodge, and W.F. Galloway, 1984,Regional controls on diagenesis and reservoir qual-ity in Lower Tertiary sandstones along the TexasGulf Coast, in D.A. McDonald and R.C. Surdam,eds., Clastic diagenesis: AAPG Memoir 37, p. 15–45.

Mark, G.M., 1978, The survivability of labile light-mineral grains in fluvial, aeolian and littoral marineenvironments: the Permian Cutler and Cedar Mesaformations, Moab, Utah: Sedimentology, v. 25, p. 587–606.

McCubbin, D.G., 1982, Barrier-island and strand plainfacies, in P.A. Scholle and D. Spearing, eds., Sand-stone depositional environments: AAPG Memoir 31,247–279.

Miall, A.D., 1984, Deltas, in R.G. Walker, ed., Faciesmodels, 2d. ed.: Geoscience Canada, Reprint Series1, p. 105–118.

Moncure, G.K., R.W. Lahann, and R.M. Siebert, 1984,Origin of secondary porosity and cement distribu-tion in a sandstone/shale sequence from the FrioFormation (Oligocene), in D.A. McDonald and R.C.Surdam, eds., Clastic diagenesis: AAPG Memoir 37,p. 151–161.

Parnell, J., 1987, Secondary porosity in hydrocarbon-bearing transgressive sandstones on an unstableLower Paleozoic continental shelf, Welch Borderland,

in J.D. Marshall, ed., Diagenesis of sedimentarysequences: Geological Society Special Publication 36,p. 297–312.

Pittman, E., 1979, Porosity, diagenesis and productivecapability of sandstone reservoirs, in P.A. Scholleand P.R. Schluger, eds., Aspects of diagenesis:SEPM Special Publication 26, p. 159–173.

Potocki, D., and I. Hutcheon, 1992, Lithology and dia-genesis of sandstones in the western Canada fore-land basin, in R.W. Macqueen and D.A. Leckie, eds.,Foreland basins and fold belts: AAPG Memoir 55, p. 229–257.

Potter, P.E., 1978, Petrology and chemistry of modernbig river sands: Journal of Geology, v. 86, p. 423–449.

Pryor, W.A., 1973, Permeability-porosity patterns andvariations in some Holocene sand bodies: AAPGBulletin, v. 57, p. 162–189.

Reinson, G.E, 1984, Barrier island and associatedstrand-plain systems, in R.G. Walker, ed., Faciesmodels, 2d ed.: Geoscience Canada, Reprint Series1, p. 119–140.

Rittenhouse, G., 1971, Mechanical compaction of sandscontaining different percentages of ductile grains: atheoretical approach: AAPG Bulletin, v. 55, p. 92–96.

Schmidt, V., and D.A. McDonald, 1979a, The role of sec-ondary porosity in the course of sandstone diagene-sis, in P.A. Scholle and P.R. Schluger, eds., Aspects ofdiagenesis: SEPM Special Publication 26, p. 175–207.

Schmidt, V., and D.A. McDonald, 1979b, Texture andrecognition of secondary porosity in sandstones, inP.A. Scholle and P.R. Schluger, eds., Aspects of dia-genesis: SEPM Special Publication 26, p. 209–225.

Sclater, J.G., and P.A.F. Christie, 1980, Continentalstretching: an exploration of the post-Mid-Creta-ceous subsidence of the central North Sea basin:Journal of Geophysics Research, v. 85, p. 3711–3739.

Shanmugam, G., 1985, Significance of secondaryporosity in interpreting sandstone composition:AAPG Bulletin, v. 69, p. 378–384.

Shanmugam, G., 1990, Porosity prediction in sandstoneusing erosional unconformities, in I.D. Meshri and P.J.Ortoleva, eds., Prediction of reservoir quality throughchemical modeling: AAPG Memoir 49, p. 1–23.

Siebert, R.M., G.K. Moncure, and R.W. Lahann, 1984,A theory of framework grain dissolution in sand-stones, in D.A. McDonald and R.C. Surdam, eds.,Clastic diagenesis: AAPG Memoir 37, p. 163–175.

Smosna, R., 1989, Compaction law for Cretaceoussandstones of Alaska’s North Slope: Journal of Sed-imentary Petrology, v. 59, p. 572–584.

Stonecipher, S.A., and J.A. May, 1990, Facies controlon early diagenesis: Wilcox Group, Texas GulfCoast, in I.D. Meshri and P.J. Ortoleva, eds., Predic-tion of reservoir quality through chemical model-ling: AAPG Memoir 44, p. 24–44.

Page 276: Reservoir Quality Prediction in Sand and Carbonates

Stonecipher, S.A., R.D. Winn, and M.G. Bishop, 1984,Diagenesis of the Frontier Formation, Moxa arch: afunction of sandstone geometry, texture and com-position, and fluid flux, in D.A. McDonald and R.C.Surdam, eds., Clastic diagenesis: AAPG Memoir 37,p. 289–316.

Swift, D.J.P., 1975, Barrier-island genesis: evidencefrom the central Atlantic shelf, eastern U.S.A.: Sedi-mentary Geology, v. 14, p. 1–43.

Walker, R.G., 1984, Shelf and shallow marine sand-stones, in R.G. Walker, ed., Facies models, 2d edition:Geoscience Canada, Reprint Series 1, p. 141–170.

Depositional Controls Over Porosity Development in Lithic Sandstones of the Appalachian Basin 265

Page 277: Reservoir Quality Prediction in Sand and Carbonates

267

Chapter 17

Predicting Reservoir Properties in Dolomites:Upper Devonian Leduc Buildups,

Deep Alberta BasinEric W. Mountjoy

Xiomara M. Marquez1

Department of Earth and Planetary Sciences, McGill UniversityMontreal, Canada

ABSTRACT

Completely dolomitized Upper Devonian Leduc buildups at depths >4000 mhave higher porosities and permeabilities than adjacent limestone buildups;dolostones are more resistant to pressure solution and tend to retain theirporosity during burial. Distribution of pore types is controlled by deposi-tional facies, whereas distribution of permeability is largely controlled bydiagenetic processes, especially dolomitization.

In pool D3A of the Strachan reservoir, porosities and permeabilities arehighest in the interior of the buildup where the strata are completely dolomi-tized. In the reef margin, porous and permeable dolomitized zones areinterbedded with nonporous and nonpermeable limestone units. The pres-ence of porous and permeable zones is closely related to the degree ofdolomitization, with the greatest porosity and permeability occurring incompletely dolomitized rocks.

The reservoir character in the Ricinus West buildup closely follows depo-sitional units, despite complete dolomitization. At the reservoir scale, porosi-ty and permeability have relatively similar values throughout the buildup.At the meter to tens of meters scale, the upper buildup interior is character-ized by 1- to 2-m-thick, permeable and laterally continuous lagoonal strata.The lower reef interior consists of laterally discontinuous permeable zones.In the reef margin, permeability is controlled by fractures and interconnectedvugs. At the millimeter scale, porosity and permeability are controlled bydiagenetic processes.

Late cementation and dissolution processes have slightly decreased andincreased porosity and permeability, mainly in the lower part of the reser-voirs. Bitumen plugging decreased porosity and permeability in the upperpart of the reservoirs. Although it is difficult to predict reservoir porosityand permeability trends, the secondary porosities in these deeply buried

1Department de Exploracion, Maraven S.A., Apartado 829, Caracas 1010-A Venezuela.

Mountjoy, E.W., and X.M. Marquez, 1997, Predictingreservoir properties in dolomites: Upper DevonianLeduc buildups, Deep Alberta Basin, in J.A.Kupecz, J. Gluyas, and S. Bloch, eds., Reservoirquality prediction in sandstones and carbonates:AAPG Memoir 69, p. 267–306.

Page 278: Reservoir Quality Prediction in Sand and Carbonates

268 Mountjoy and Marquez

INTRODUCTION

The objective of this study is to better understanddeeply buried limestone and dolomite reservoirs sothat variations in reservoir porosity and permeabilitycan be predicted in advance of drilling. Depositionalfacies and textures, diagenetic alteration, and burialeffects are among the factors that can contribute toreservoir heterogeneity. We use examples from thedeep basin of the Western Canada Sedimentary Basinas the focus of our study because they constitute pro-lific and widespread petroleum reservoirs.

Except for facies descriptions and the interpretationof depositional environments, few carbonate reservoirsin the Upper Devonian of the Western Canada Sedi-mentary Basin have been studied in detail for reservoircharacter. Leduc buildups (Upper Devonian, Frasnian)along the 320-km-long Rimbey-Meadowbrook reeftrend in central Alberta (Figure 1) are prolific oil and gasproducers. For example, the Ricinus West pooldiscovered in 1969 has produced 825 bcf, and the Stra-chan D3A pool discovered in 1967 has produced 911bcf, with estimated original gas in place of 1.8 tcf and1.4 tcf, respectively, for each field. The H2S contentvaries from 3.2% to 10.8% in the Strachan D3A pool to25% in the D3B pool, and 31% to 33.5% in RicinusWest. The Ricinus West field is larger (8 × 4 km) thanStrachan D3A (6 × 3 km). The cumulative productioncompared to single section reserves of gas in placesuggests that the dolomitized wells in Ricinus Westare draining considerably more than one section andindicates that there is good reservoir continuity in thisfield (Podrusky et al., 1987).

The stratigraphy and geological setting of theRimbey-Meadowbrook buildups have been outlinedin Amthor et al. (1993, 1994), and their regional settingand depositional history have been described byStoakes (1992). Information concerning the facies anddepositional environments of Leduc buildups ismainly available from the undolomitized Redwaterand Golden Spike buildups (Klovan, 1964; Wendte,1974; McGillivary and Mountjoy, 1975; Reitzel et al.,1976; Burrowes, 1977; Jardine et al., 1977; Reitzel andCallow, 1977; Walls, 1978; Walls et al., 1979; Jardineand Wishart, 1982; Walls and Burrowes, 1985; Bur-rowes and Krouse, 1987; Carpenter and Lohmann,1989; Wendte 1992a, b, 1994; Chouinard, 1993). A few

studies discuss the extensively dolomitized part of thereef trend (Figure 1) (Layer, 1949; Waring and Layer,1950; Andrichuk, 1958a, b; Illing, 1959; Barfoot andRodgers, 1984; Barfoot and Ko, 1987; Machel andMountjoy, 1987; McNamara and Wardlaw, 1991;Amthor et al., 1993; Drivet, 1993; Drivet and Mountjoy,1993, 1997; Amthor et al., 1994; Marquez, 1994).

In terms of reservoir character in limestone buildups,only the Leduc Golden Spike (McGillivary andMountjoy, 1975; Walls, 1978, 1983; Walls and Burrowes,1985) and Swan Hills Judy Creek limestone reservoirshave been studied in detail (Wendte and Stoakes, 1982;Walls and Burrowes, 1990). In the case of dolomite reser-voirs, only the Westerose buildup of the Rimbey-Meadowbrook reef trend has been studied (McNamaraand Wardlaw, 1991). Mountjoy (1994) summarized thecharacter of dolomitized reservoirs of the Devonian ofwestern Canada. The present study and that of Drivet(1993) were designed to investigate the reservoir charac-ter of dolomite reservoirs in order to provide a suitabledatabase that could be used for the prediction of reser-voir quality. Not only was diagenesis and its modifica-tion of primary facies and porosity examined, but alsothe effects of diagenesis on reservoir quality withincreasing depth. Distinguishing primary porosity andpermeability from textures and fabrics that are over-printed by diagenesis and dolomitization (Mazzulo,1992; Mountjoy, 1994) is difficult because primary faciesand textures are often greatly modified or destroyed.

This chapter summarizes the variability of reservoirquality (porosity and permeability) in deeply buriedcarbonates, and addresses the potential controls ofdepositional facies and diagenesis on pore systems.Reservoir character was studied at three differentscales: the entire reservoir, meter-scale depositionalunits, and the individual pore types (Weber, 1986). Theobjectives are: (1) to identify different pore types and todetermine their distribution within Leduc reservoirs;(2) to determine reservoir continuity and variability,and porosity and permeability trends, relative to differ-ent depositional facies and diagenetic phases; (3) tocompare reservoir characteristics of limestone anddolostone reservoirs in the deep basin; and (4) to deter-mine how dolomitization, cementation, and reservoirbitumen have affected reservoir character.

dolomites are mainly controlled by the primary porosity distribution and thedepositional facies. The permeability is mainly controlled by diageneticprocesses, especially dolomitization and various phases of cementation andbitumen plugging in the upper part of the reservoirs. Available data fromthe deep basin and the adjacent Rocky Mountains suggest that these porousdolomites are regionally extensive, and dolomite buildups elsewhere shouldhave porosity and permeability variations similar to the Strachan andRicinus West reservoirs. However, late-stage dolomite, anhydrite, and bitu-men can locally partially to completely fill the pore spaces.

Page 279: Reservoir Quality Prediction in Sand and Carbonates

The reservoir characteristics of these Leducbuildups have been partly documented in the shallow-(<2000 m) and intermediate-burial (2000 to ~3500 m)portion of the Rimbey-Meadowbrook reef trend(Reitzel and Callow, 1977; Barfoot and Rodgers, 1984;Barfoot and Ko, 1987; Hugo, 1990; McNamara andWardlaw, 1991; Drivet, 1993; Amthor et al., 1994; Dri-vet and Mountjoy, 1994, 1997). In the deeper (>4000 m)part of the reef trend, there are few published data onreservoir properties and their variability within theLeduc buildups, except for a summary of regional-scale porosity and permeability variations by Amthoret al. (1994) and brief reports on the Strachan and Rici-nus West gas reservoirs by Hriskevich et al. (1980) andSeifert (1990). Some information is also available in

theses (Laflamme, 1990; Drivet, 1993; Marquez, 1994).Consequently, little is known about the porosity andpermeability variations within these Leduc reservoirs.The partly dolomitized Strachan and completelydolomitized Ricinus West buildups provide an idealarea for comparing the reservoir characteristics ofdeeply buried limestone and dolostone buildups.Information on the reservoir characteristics of theintermediately buried (2500 m) Homeglen-Rimbeydolomitized buildup is available in Drivet (1993) andDrivet and Mountjoy (1997).

On a regional scale, at shallow depth (<2000 m),porosity and permeability of limestones and dolo-stones have comparable values and distributions ofporosity (Amthor et al., 1994). At these depths,

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 269

��������������������

���������

���������

��

����

���

�������

��

T55

��Golden Spike

Edmonton

R1 R27 R20

5th

T50

T45

T50

T45

T40

Western Margin ofCooking Lake Platform

��

Extent of dolomitizedCooking Lake

Morinville

(underlain by Cooking Lake Platform)

������

����

���������

��

������������

����������

����������

Bonnie Glen

Westerose

Bearberry

T35

Wizard Lake

Leduc

Acheson

Redwater

Homeglen Rimbey

20 km

�?

������

Ricinus .

W

E

�����Strachan

Medicine RiverSTUDY AREA

Harmattan

West Ireton Shale BasinEast Ireton Shale Basin

R10

��������������������������������������������������������

���������

����������������������������

RICINU

S W

CHEDDERVILLE

EDG

E OF TH

RU

ST-FOLD

BELT

�������������������������������������������

CRIMSON

STRACHAN

BEARBERRY

RICINU

S E.

20 km

T35

R10 R8

��������

���� PHOENIX

R5

6-36

10-33

11-27 7-266-25

7-13 6-10

11-17

6-24

6-14

10-24

15-23

���������

��������D3A

D3B

5-4

4-15

13-9

7-19

7-9

Figure 1. Study area in thedeep basin at the southernend of the Rimbey-Meadowbrook reef trend inthe center of the AlbertaBasin. Inset map shows thedistribution of the Strachan,Ricinus West, and Eastfields, and adjacentbuildups.

Page 280: Reservoir Quality Prediction in Sand and Carbonates

270 Mountjoy and Marquez

dolomitization resulted in the redistribution and somedecrease of porosity and a slight increase in perme-ability (Amthor et al., 1994). At deeper burial (>2000m), however, dolostones are significantly more porousand permeable than adjacent limestones (Amthor etal., 1994; Drivet and Mountjoy, 1994; Mountjoy andAmthor, 1994), as also occurs in South Florida(Schmoker and Halley, 1982).

METHODS

All available cores from the Strachan and RicinusWest buildups in the Upper Devonian of centralAlberta, between townships 34 to 39 and ranges7–12W5 (Figures 2–5; Table 1) were logged and sam-pled from depths of ~3500 to 5000 m. Core parameterswere observed and recorded systematically (Table 2).Core-derived horizontal permeabilities (Kh) and verti-cal permeabilities (Kv) and porosities were obtainedfrom the Energy Resources Conservation Board(ERCB), Calgary. Permeability and porosity data fromthe gas-producing zone were measured from whole-core samples (full diameter, 8 cm; 0.3 m long). Becausesample length is large relative to width, a potentialbias exists because the horizontal to vertical ratio is toolarge (Lishman, 1969). Reservoirs can be consideredstratified with respect to permeability because of verti-cal facies changes. Statistical analyses were obtainedusing the software STATVIEW II (1990). Permeability val-ues >2000 md were excluded because they probablyrepresent unrealistic values due to fracturing or largevugs. A few logs were analyzed using GEOGRAPHIXQLA2 log analysis software.

To compare porosity among different facies, an arith-metic mean of the core measurements is used, becauseporosity is a scalar quantity and commonly normallydistributed. Permeability has a log normal or skeweddistribution and commonly is highly variable overshort distances, making a geometric mean a moreappropriate choice (Wardlaw, 1990, 1992). McNamaraand Wardlaw (1991) reported that the geometric meanof core-measured permeability provided the best corre-spondence, with permeability estimated from a pres-sure buildup test in the Leduc Westerose reservoirupdip along the Rimbey-Meadowbrook reef trend (Fig-ure 1). Permeability and porosity profiles within dolo-stones were obtained by plotting porosity, Kh, and Kvvalues for each sample depth for most of the wells stud-ied in the Ricinus buildup (Marquez, 1994), with repre-sentative portions illustrated in this study. Arithmetic,geometric, and harmonic means of the permeabilitywere calculated using the software STATVIEW II (1990).

Information from core observations and their corre-lation with porosity and permeability data are empha-sized in this study for reservoir quality prediction.Maddox (1984) determined the correlation betweencore porosity and log porosity from neutron, density,and acoustic logs over the cored interval in the RicinusWest reservoir. Core analyses yielded similar toslightly higher porosities compared with those calcu-lated from logs. Comparisons of porosities from cores

and logs of Leduc dolomites along the Rimbey-Mead-owbrook reef trend made by McNamara et al. (1991)and Drivet (1993) indicate that core porosities are gen-erally comparable to porosities calculated from logs,when suitable logs are available. Furthermore, McNa-mara et al. (1991) showed that when a core containspores larger than the core diameter of 8 cm, the valuemeasured directly from cores underestimated theporosity by 3% or more. Thus, core analysis measure-ments can be considered reliable for the recognition ofporosity and permeability trends, but will tend tounderestimate porosity in coarse, vuggy carbonates.

Measurements of permeability at simulated reser-voir conditions are one order of magnitude lower thansimilar measurements at ambient pressure (Vavra etal., 1991), so that measured values (ambient) representmaximum permeability. Which geologic parametersshould be described and mapped to permit a reason-ably accurate petrophysical quantification of carbon-ate reservoir models has been discussed and outlinedby Lucia (1983, 1995), and Lucia and Conti (1987).Lucia (1995) advocates using rock-fabric units basedon grain size and sorting, dolomite crystal size, sepa-rate-vug type and porosity, and total porosity.

FACIES AND DIAGENESIS

The Leduc Formation of the Strachan (Figures 2, 3)and Ricinus West fields (Figures 4, 5) are stromato-poroid-coral buildups of Upper Devonian age (LeducFormation) with generally similar depositional facies(Marquez, 1994). The buildups consist of reef marginand reef interior environments (summarized in Figures3, 5). The reef margin facies include coral rudstones,tabular stromatoporoid boundstones, and stromato-poroid-coral rudstones (Figure 6), indicating depositionin shallow, high-energy environments. The buildupinterior comprises skeletal packstones/grainstones,skeletal wackestones, microbial laminites, and locallygreen shales (Figure 7) suggestive of shallowing-upward, peritidal conditions. The interior of the RicinusWest buildup is divisible into lower and upper parts.The lower portion consists of domal stromatoporoidfloatstones, skeletal wackestones, and coral rudstones.The upper part (131 m) consists of six shallowing-upward parasequences (generally 8 m thick, but ≤27 m)and comprise from bottom to top: Amphipora-richwackestones or packstones, locally with small domalstromatoporoids; skeletal grainstones; and peritidalmicrobial laminites. These cycles are two to three timesas thick as outcrop equivalents (McLean and Mountjoy,1994), which may be due to obliteration of some faciesby dolomitization, or continuous deposition withoutobvious breaks (Marquez, 1994).

The limestone Strachan and dolomitized RicinusWest fields have undergone a complex diagenetic his-tory and different diagenetic overprints (Figure 8) (Mar-quez, 1994). The Ricinus West buildup is similar to otherdolomitized Leduc buildups along the Rimbey-Meadowbrook reef trend (Amthor et al., 1993, 1994;Drivet and Mountjoy, 1996). Near-surface sea-floor

Page 281: Reservoir Quality Prediction in Sand and Carbonates

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 271

300

0

500

300

0

500

700

600

300

0

R. 1

0R

.9W

5

TWP 38TWP 37

G/W CONTACTAT -9663 ft (-2946 m)

D3A POOL

G/W CONTACTAT -9560 ft (-2915 m)

D3B POOL

10-31

15-2

5-11

14-2

7-32

11-27B

miles

kilometers

0

0

1 2

1 2 3

CORED WELL

DRY HOLE

GAS WELL

A

A'

B'

B

NW

SE

D3A Pool

D3B Pool

12-31

11-27A

11-22

������

��������������������

���������

���������

����

����

���

����

���

��

����

Pool D3B

COOKING LAKE EQUIVALENT

BEAVERHILL LAKE

50

0

m

km

0 1 2

9560' G/W

9663' G/W

14-2

5-11 15-2 12-31

10-31

7-32

11-27

11-22

SENW

10,613 ft.

Ireton

Calc. Shales& Argil. Limestones

Pool D3A

??

Nisku

Ireton core

Iretoncore

Duvernay?

Leduc FormationLimestone anddolomite

Figure 2. Leduc net pay isopach map of the Strachan buildup. (A) Map of well locations and netpay thickness (modified from Hriskevich et al., 1980). Wells 5-11, 14-2, 15-2, and 12-31 are non-productive. (B) A NW-SE cross section shows distribution of limestone and partly dolomitizedlimestone (white), approximate gas/water (G/W) contact, and distribution of core.

Page 282: Reservoir Quality Prediction in Sand and Carbonates

272 Mountjoy and Marquez

SWCRSCR

SCR

TSB

SF SF SFSGSG

5-11

-38-

10W

5

SWSW

Top

of C

alm

ar

SC

R

SW

SW

CR

SC

R

SC

R

TS

B

SC

R

CR

*

***

**

SWS

kele

tal W

acke

ston

es

Dom

al S

trom

atop

oroi

dF

loat

ston

esSF

Cor

al R

udst

ones

CR

Tab

ular

Stro

mat

opor

oid

Bou

ndst

ones

TSB

Str

omat

op.-

Cor

al R

udst

ones

SCR

SP/G

Ske

leta

l Pac

ksto

nes/

Gra

inst

ones

No

core

rec

over

ed

O(%

)K

h (m

d)Kv

(m

d)

*

SCR

SW

SCR

SCR

CR

TSB

TSB

TSB

SW**

SW

SW SP

/G

O(%

)K

h (m

d)Kv

(m

d)

SW11-2

2-37

-9W

5

*

SW SW SW SW SW

**

**

CR

SW

Top

of

Led

uc F

m.

T

op o

fL

educ

Fm

.

BPO

OL

D3A

124

m

185

m

Top

of C

alm

ar

12-3

1-37

-9W

3

CR

0-10

0-20

0-3

SCR

SCR

Kh

Kv

Iret

on F

m.

99 m

114

mB'

156

m

10-3

1-37

-9W

5

AA

'

175

m

134

m13

7 m

7-32

-9W

5

11-2

7-37

-9W

5

No Recovery

14-2

-38-

10-W

515

-2-3

8-10

W5

* Pe

rmea

bilit

y be

twee

n

0.01

and

1 m

d**

Perm

eabi

lity

betw

een

0.

01 a

nd 0

.3 m

d

SCA

LE15

m

REE

F

REE

F M

AR

GIN

Stra

chan

Bui

ldup

DEP

OSI

TIO

NA

L FA

CIE

S

POO

L D

3B

Iret

on F

m.

Figu

re 3

. Dis

trib

uti

on o

f d

epos

itio

nal

fac

ies

in S

trac

han

D3A

(sec

tion

BB

’) a

nd

D3B

poo

ls (s

ecti

on A

A’)

. See

Fig

ure

2 f

or w

ell l

ocat

ion

s. D

atu

mis

top

of

Cal

mar

For

mat

ion

. To

the

righ

t of

each

col

um

n a

re a

rith

met

ic a

vera

ge p

oros

ity

in p

erce

nt a

nd

geo

met

ric

aver

age

per

mea

bil

itie

s in

mil

lid

arci

es. P

oros

ity

and

per

mea

bil

ity

dat

a ar

e b

ased

on

cor

e an

alys

es f

rom

the

En

ergy

Res

ourc

es C

onse

rvat

ion

Boa

rd (E

RC

B) i

n C

alga

ry.

Page 283: Reservoir Quality Prediction in Sand and Carbonates

radiaxial fibrous calcites fill intergranular, shelter, andgrowth framework cavities in the buildup margins. Syn-taxial overgrowths are rare and were followed by theprecipitation of blocky calcite cements in skeletal pores.During intermediate burial, chemical compaction andpartial to complete replacement dolomitizationoccurred. Complete dolomitization in the Ricinus Westbuildup obliterated many of the early paragenetic fea-

tures. Most of the Leduc carbonates appear to have beenreplaced by dolomites at temperatures of ~40°–60°C anddepths of 500–1200 m during the Late Devonian to EarlyMississippian (Amthor et al., 1993; Mountjoy andAmthor, 1994; Drivet and Mountjoy, 1997). The sourcesand flow directions of the dolomitizing fluids are stilluncertain. Replacement dolomitization was followed byfracturing, dissolution, and minor dolomite cement.

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 273

������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������

���

���

���

���

��������

������

���

UPPER LEDUCMIDDLE LEDUC

�����������������������������������

NW SE

(?)

B)

��������������������

100

0

0

m

Ireton Fm.

10-33 11-27

21

7-26

15-23

6-25

7-13A)

LEDUC Fm.

DUVERNAY Fm.DUVERNAY Fm.

COOKING LAKE Fm.

km

TWP 37TWP 36

R. 1

0R

. 9W

.5

0

400

0

600

0

500

0300

300

G/W CONTACT -10,500 ft(-3201 m)

11-27

7-26 6-25

7-13

kilometers

0

0

1 2

1 2 3

miles

CORED WELL

GAS WELL

15-23

NW

SE

B

B'

A

A'

10-33

Figure 4. RicinusWest buildup. (A) A NW-SE crosssection showingstratigraphy andlocation of cores.(B) Map of welllocations and netpay thickness(modified fromHriskevich et al.,1980). G/W =gas/water.

Page 284: Reservoir Quality Prediction in Sand and Carbonates

274 Mountjoy and Marquez

AL

SP/G

SP/G

SP/G

L

SW

SW

SW

SW

L

SP/G

SP/G

SP/G

L

SP/G

Ø Kh

0 - 10 0 - 40

7-26L

Calmar topBUILDUP INTERIOR

A

SW

SP/G

L

L

SF/R

CR

SW

UPPER LEDUCMIDDLE LEDUC

?

KvKh0 - 10 0 - 40 0 - 6

KvKh

11-27

0-10 0-40 0-6Ø

Ø

UP

PE

R

BU

ILD

UP

I

NT

ER

IOR

LO

WE

R

BU

ILD

UP

IN

TE

RIO

R

Fig

. 1

1F

ig.

14

Laminates

Skeletal Wackestones

Skeletal Packstones/Grainstones

Green Shales

Stromatoporoid Floatstones-

Rudstones

Coral Rudstones

Tabular Stromatoporoid

Rudstones

Coral Wackestones

CR

SG

SF/R

SP/W

SP/W

SP/W

SG

L

CW

L

Calmar top

Leduc top

BUILDUP MARGIN

DEPOSITIONAL FACIES

Ø KvKh

0 - 10 0 - 40 0 - 6

Kh Kv

0 - 10 0 - 40

B B'

15-23

7-13

Ø

10-33

5 km

Ireton Fm.98.4 m

Leduc top

99.6 m

62.5 m

LSW

SP/G

SF/R

CR

TSB

CW

2.5 km

Ireton Fm. 99 m

SG

SG

159 m

0 - 6

Kv

TSB

SG

SG

TSB

SG

TSB

TSB

SG

SG

SG

SF/R

CR

CR

7-13

15 m

0 - 6

BU

ILD

UP

MA

RG

INF

ig.

12

BU

ILD

UP

MA

RG

IN

Figure 5. Depositional facies in the Ricinus West buildup interior wells 10-33, 11-27, and 7-26, and margin wells(lower right) 15-23 and 7-13 (see Figure 4 for well locations). Datum is top of Calmar Formation. To the right ofeach column are the arithmetic average porosity in percent and geometric average horizontal (Kh) and vertical(Kv) permeabilities in millidarcies. Porosity and permeability data are based on core analyses from ERCB.

Page 285: Reservoir Quality Prediction in Sand and Carbonates

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 275

Tab

le 1

. Wel

ls a

nd

Cor

ed I

nte

rval

s fr

om th

e S

trac

han

, Ric

inu

s, a

nd

Ad

jace

nt B

uil

du

ps.

Wel

lL

ocat

ion

Cor

ed I

nte

rval

(ft)

Cor

ed I

nte

rval

(m)

Tot

al (f

t)T

otal

(m)

Fiel

d

Ban

ff A

quit

ane

10-3

1-37

-9W

513

,442

–14,

200

4098

.2–4

329.

375

8.0

231.

1St

rach

anB

anff

Aqu

itan

e12

-31-

37-9

W5

14,0

46–1

4,09

642

82.3

–429

7.6

50.0

15.2

Stra

chan

Aqu

itan

e15

-2-3

8-10

W5

13,4

85–1

3,60

541

11.3

–414

7.9

120.

036

.6St

rach

anC

hevr

on S

OB

C14

-2-3

8-10

W5

13,4

66–1

3,52

641

05.5

–412

3.8

60.0

18.3

Stra

chan

Aqu

it. S

hell-

Stra

chan

5-11

-38-

10W

513

,916

–13,

702

4242

.7–4

177.

424

5.0

74.7

Stra

chan

Ban

ff A

quit

ane

7-32

-37-

9W5

13,4

22–1

3,49

440

92.1

–411

4.0

72.0

22.0

Stra

chan

B.A

. et a

l.11

-27-

37-9

W5

13,1

87–1

3,19

740

20.4

–402

3.5

10.0

3.0

Stra

chan

Gul

f et a

l. St

rach

an11

-22-

37-9

W5

12,9

70–1

3,04

639

54.3

–399

7.4

76.0

23.2

Stra

chan

Aqu

itan

e et

al.

10-1

6-37

-10W

515

,306

–15,

333

4666

.5–4

674.

727

.08.

2St

rach

anG

ulf e

t al.

Stra

chan

10-2

4-37

-9W

513

,322

–13,

378

4061

.6–4

078.

756

.017

.1C

rim

son

Arc

ho P

acif

ic F

ina

Cow

Jk7-

33-3

7-8W

512

,133

–12,

168

3699

.1–3

709.

835

.010

.7C

rim

son

Hus

ky e

t al.

11-2

8-37

-8W

512

,133

–12,

178

3699

.1–3

712.

845

.013

.7C

rim

son

Gul

f Poc

et a

l.7-

19-3

7-8W

513

,436

–13,

493

4096

.3–4

113.

757

.017

.4C

rim

son

Imp.

Che

dd

ervi

lle10

-20-

37-7

W5

11,6

08–1

1,68

735

39.0

–356

3.1

79.0

24.1

Che

dd

ervi

llePi

nn. C

hed

der

ville

16-1

9-37

-7W

511

,642

–11,

692

3549

.4–3

564.

650

.015

.2C

hed

der

ville

Imp.

Che

dd

ervi

lle10

-29-

37-7

W5

11,9

91–1

2,01

636

55.8

–366

3.4

25.0

7.6

Che

dd

ervi

lleB

P C

hed

der

ville

6-30

-37-

7W5

11,5

85–1

1,59

535

32.0

–353

5.1

9.8

3.0

Che

dd

ervi

lleD

ome

et a

l. C

hed

der

ville

14-9

-37-

7W5

11,6

34–1

1,71

335

47.0

–357

1.0

79.0

24.1

Che

dd

ervi

lleE

sso

Mob

il R

icin

us10

-11-

37-8

W5

12,3

56–1

2,43

837

67.1

–379

2.1

82.0

25.0

Che

dd

ervi

lleIm

p. H

B. R

icin

us12

-28-

36-7

W5

12,4

52–1

2,51

837

96.3

–381

6.5

66.0

20.1

Che

dd

ervi

lle

Ban

ff e

t al.

10-3

3-36

-10W

514

,785

–15,

319

4507

.6–4

670.

453

4.0

162.

8R

icin

usB

anff

Aqu

itan

e11

-27-

36-1

0W5

14,6

64–1

4,72

444

70.7

–448

9.0

60.0

18.3

Ric

inus

Che

vron

Mob

il7-

26-3

6-10

W5

14,3

81–1

4,43

443

84.5

–440

0.6

53.0

16.2

Ric

inus

Ban

ff A

quit

ane

15-2

3-36

-10W

514

,616

–15,

067

4456

.1–4

593.

645

1.0

137.

5R

icin

usB

anff

et a

l.7-

13-3

6-10

W5

14,2

93–1

4,86

543

57.6

–453

2.0

572.

017

4.4

Ric

inus

Ban

ff e

t al.

Ric

inus

6-25

-36-

10W

5514

,323

–14,

378

4366

.8–4

383.

555

.016

.8R

icin

usM

obil

Ric

inus

6-10

-36-

9W5

14,2

40–1

4,25

143

41.5

–434

4.8

11.0

3.4

Ric

inus

Mob

il et

al.

Ric

inus

11-1

7-35

-8W

513

,814

–13,

921

4211

.6–4

244.

210

7.0

32.6

Ric

inus

Pan.

Am

er. R

icin

us6-

24-3

4-8W

513

,414

–13,

475

4089

.6–4

108.

261

.018

.6R

icin

usA

lban

y A

moc

o R

icin

us6-

14-3

4-8W

514

,048

–14,

138

4282

.9–4

310.

490

.027

.4R

icin

us

Ban

ff A

quit

. Ram

Riv

er7-

9-37

-10W

515

,404

–15,

510

4696

.3–4

728.

710

8.0

32.9

Ram

Riv

erH

usky

Ram

Riv

er10

-16-

37-1

0W5

15,3

06–1

5,33

346

66.5

–467

4.7

27.0

8.2

Ram

Riv

erSh

ell C

ante

rra

Ram

R.

5-13

-37-

12W

516

,628

–16,

515

5069

.5–5

035.

111

3.0

34.5

Ram

Riv

erU

no-T

ex e

t al.

Phoe

nix

5-4-

39-1

1W5

14,2

20–1

4,28

043

35.4

–435

3.7

60.0

18.3

Phoe

nix

RR

Am

oco

et a

l. A

ncon

a7-

9-39

-12W

515

,539

–15,

553

4737

.5–4

741.

814

.14.

3Ph

oeni

xC

.S. e

t al.

Phoe

nix

6-36

-38-

11W

513

,884

–13,

910

4232

.9–4

240.

926

.48.

0Ph

oeni

x

Page 286: Reservoir Quality Prediction in Sand and Carbonates

276 Mountjoy and Marquez

Anhydrite cements postdate the dolomite cements.Microfractures filled with bitumen crosscut most diage-netic features. These hairline microfractures probablyformed during overpressuring of a well-sealed reservoirduring progressive burial by thermal cracking of crudeoil to gas in conjunction with shearing related to tectoniccompression (Marquez and Mountjoy, 1996). Calcitecements postdate microfracturing, with the latest phaserelated to thermochemical sulfate reduction (Krause etal., 1988; Marquez, 1994) (Figure 8).

Of critical importance to the rock fabrics, especiallythe permeability, is the size of the grains or dolomitecrystals. Lucia (1983, 1995) grouped nonvuggy carbon-ates (both limestones and dolostones) into threeporosity-permeability fields defined using particle andcrystal-size boundaries of 20 and 100 µm. Fine tomedium crystalline grainstone-dominated dolopack-stones and medium crystalline mud-dominated dolo-stones plot in the 20–100 µm permeability field, andmore coarsely crystalline rocks generally plot above the100-µm boundary (Lucia, 1995). The replacementdolomites of the Strachan and Ricinus reservoirs consist

predominantly of two types: (1) medium crystalline(60–250 µm), planar, euhedral to subhedral, formingdense and porous mosaics; and (2) fine crystalline (30–60µm), planar, euhedral to subhedral, forming dense andporous mosaics. The medium crystalline dolomites arethe most common, forming ≤90% of some wells (e.g.,Stachan D3A 10-31, Crimson 10-24), and most wouldplot above the 100-µm boundary. The fine crystallinedolomites occur in the matrix of partially dolomitizedpackstones and rudstones (e.g., Strachan well 14-2) andwould plot within the 20- to 100-µm boundaries. A thirdtype, coarsely crystalline (250–600 µm), is locally abun-dant in Ricinus well 11-27. A fourth type, touching vugswith a matrix of intercrystalline dolomite, occurs in boththe Strachan and Ricinus reservoirs.

PORE TYPES AND DEFINITIONS

The following sections summarize the pore types,permeabilities, and the resulting porosity networkwithin these limestone and dolostone reservoirs (Tables3–6). The Strachan reservoir has a weak water drive and

Table 2. Systematic Observations Made from Core.

1. Core Number, Box, Depth, Recovery (%)

2. Lithologya. Limestones

Texture, fossils, contactsDepositional faciesDegree of dolomitization

b. DolostonesCrystal SizeFine: 30–62 µmMedium: 62–250 µmVery Coarse: >600 µm

3. Pore TypeIntraskeletal

Fossil typeIntercrystallineMoldic

AmphiporaThamnopora-like

VugFractureBreccia

4. Pore SizeVery large: >1.0 cmLarge: 0.5–1.0 cmMedium: 2.0–5.0 mmSmall: 0.1–1.0 mmVery small: >0.1 mm

5. Pore ShapeSphericalTubularTabularIrregularPolyhedral

6. Pore Association ConnectionMatrix (intercrystalline)FracturesTouching vugs/molds

7. CementsType

DolomiteAnhydriteSulfidesSulfurQuartzReservoir bitumenCalcite

Pore TypeDegree of Filling

OpenPartly filledFilled

8. FracturesOrientation

SubverticalSubhorizontal

IntensityWidth, LengthFilling

9. BrecciasCrackleMosaicRubble

Page 287: Reservoir Quality Prediction in Sand and Carbonates

is thus a partially “closed” system that underwent rela-tively rapid pressure decline during production (Hriske-vich et al., 1980). A good understanding of the poretypes and their distribution is essential for efficientrecovery of the hydrocarbons in these reservoirs.

Pores types are characterized by different shape,size, orientation, and interconnection (Tables 3, 4), andare classified following Choquette and Pray (1970).They form three broad groupings with respect to their

distribution within the buildup. Dense matrix refers tocompact crystalline texture (Archie, 1952). Matrix inter-crystalline porosity is defined as the spaces betweendolomite crystals, except where there is evidence of dis-solution, and larger pores (small vugs) are present(equivalent to the “pinpoint” porosity of McNamaraand Wardlaw, 1991). Fracture porosity is subdividedinto unfilled, partly filled, and filled, because fracturesthat appear to be filled may have permeability that is

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 277

Figure 6. Core photographs showing distibution of dolomite in Strachan buildup. Scales for cores are in centime-ters. (A) Selective replacement of lime mud matrix by isolated rhombs (110 µm) and patches (arrows) of dolomitesin stromatoporoid (S) and coral (C) rudstones facies. Minor intercrystalline porosity, D3A 10-31-37-9W5, 4281 m. (B) Selective replacement of matrix and initiation of dolomite patches and mosaics(light gray). Compare with completely dolomitized sample (6F) only 0.5 m apart. Stromatoporoid (S) and coralrudstones facies, D3A 10-31-37-9W5, 4279.8 m. (C) Calcite marine cements (mc) and skeletal fragments surroundedby completely dolomitized matrix. Stromatoporoid and coral rudstones facies, D3A 10-31-37-9W5, 4181.4 m. (D) Replacement dolomite completely replaces matrix (light gray). Intraskeletal pores in stromatoporoid fragment(S) are partly filled with reservoir bitumen. Small vugs occur in the dolomitized matrix. Stromatoporoid and coralrudstones facies, D3A 10-31-37-9W5, 4265 m. (E) Replacement dolomitization of matrix results in a nonporousmosaic of dolomite crystals with some vugs. Partly leached domal stromatoporoid fragments. Reef marginstromatoporoid and coral rudstones facies, D3A 12-31-37-9W5, 4287.5 m. (F) Dissolution of skeletal fragmentsresults in molds, solution-enlarged molds, and vugs. Pores are lined by reservoir bitumen. Stromatoporoid andcoral rudstone facies, D3A 10-31-37-9W5, 4279.2 m.

Page 288: Reservoir Quality Prediction in Sand and Carbonates

278 Mountjoy and Marquez

Figure 7. Core photographs showing depositional facies and associated pore types in dolomitized Ricinus Westbuildup interior well 10-33-36-10W5. Scales for cores are in centimeters. (A) Skeletal packstone facies, leachedAmphipora moldic pores in a tight packstone matrix; 4601.5 m. (B) Skeletal wackestone facies, isolated tubular,moldic Amphipora pores (arrows) in a tight mudstone matrix; 4605.4 m. (C) Skeletal wackestone facies: intercrys-talline pores partly filled with reservoir bitumen (black/staining), small isolated moldic pores (arrows), andvugs; 4601.2 m. (D) Skeletal wackestone facies: patches (P) of polyhedral intercrystalline pores partly filled withbitumen in an otherwise tight matrix. Stromatoporoid fragment (bottom) with solution-enlarged, tabularintraskeletal pores (S). Subvertical fracture (F) is partly filled with late calcite cement; 4579.2 m. (E) Microbiallaminite facies: finely laminated mudstones with elongated, irregular, fenestral-like pores; 4608.5 m. (F)Stylolitic contact between microbial laminite facies and mudstones with pores filled with green shales; 4576.2 m.(G) Domal stromatoporoid floatstone facies: partly leached domal stromatoporoid (DS) with intraskeletal porespartly filled with reservoir bitumen (black). Late calcites (Ca) completely fill the remaining pore space; 4627.4 m.(H) Domal stromatoporoid rudstones: partly leached stromatoporoid (S) and coral (arrow) fragments in adolomudstone matrix; 4664 m. (I) Green shaly laminations (arrow) and associated stylolites. Bottom half of coreshows tight dolomudstone with vug completely filled with anhydrite; 4632.6 m.

Page 289: Reservoir Quality Prediction in Sand and Carbonates

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 279

DIAGENETICEVENT

OLDEST YOUNGEST

Micritization

Radiaxial calciteCementation

Dissolution

Minor silicaReplacement

Syntaxial calciteCements

Blocky calciteCements

Stylolitization

Fracturing

Dissolution

Dolomite Cements

BitumenEmplacement

Late calciteCements

Shallow Intermediate DeepSea floor

??

Replacement-typeDolomite rhombs

HairlineMicrofracturing

Dissolution

?

DIAGENETICEVENT

OLDEST YOUNGEST

Stylolitization

ReplacementDolomitization(R1 to R3)

Dissolution

Minor DolomiteCements

BitumenEmplacement

Late calciteCements

Shallow Intermediate Deep

HairlineMicrofracturing

Minor AnhydriteReplacement andCementation

? ?

Dissolution

Early diagenesis Obliterated by Dolomitization

Minor PyriteSphaleriteEmplacement

? ?

2

1

1

B)

A)

Figure 8. Paragenetic sequence for Strachan buildup: (A) limestone portion and (B) completely dolomitized portion, which also applies to Ricinus West and adja-cent buildups. The onset of stylolitization (>500 m) designates the beginning ofintermediate burial, and maturation of organic matter designates deep burial.

Page 290: Reservoir Quality Prediction in Sand and Carbonates

280 Mountjoy and Marquez

Tab

le 3

. Por

e T

ypes

in L

imes

ton

e an

d P

artl

y D

olom

itiz

ed S

trat

a.*

Page 291: Reservoir Quality Prediction in Sand and Carbonates

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 281

Tab

le 4

. Por

e T

ypes

in D

olos

ton

es.

Page 292: Reservoir Quality Prediction in Sand and Carbonates

282 Mountjoy and Marquez

Table 5. Porosity and Permeability Values in Limestones, Partly Dolomitized Rocks, and Dolostones.*

Porosity Kh (md) Kv (md)Depositional Facies Ar Min Max Geo Min Max Geo Min Max N

LIMESTONESSTRACHAN BUILDUP MARGINStromatoporoid Floatstone N/A N/A N/ASkeletal Wackestones 3.6 0.5 7.1 0.8 0.01 16.2 0.4 0.01 6.4 82Tabular Stromatoporoid Boundstone 2.3 0.3 8.6 0.1 0.01 360 0.02 0.01 6.5 92Stromatoporoid Coral Rudstone 2.6 0.3 15.4 0.1 0.01 2.1 0.01 0.01 74 179Coral Rudstone 3.0 0.4 29.0 0.3 0.01 247 0.01 0.02 55 95

PARTLY DOLOMITIZEDSTRACHAN BUILDUP MARGINStromatoporoid Floatstone N/A N/A N/ASkeletal Wackestones 4.91 2.0 8.0 0.21 0.01 20.7 0.02 0.01 11.5 8Tabular Stromatoporoid Boundstone 3.5 0.3 9.2 1.1 0.01 715 0.1 0.01 9.8 74Stromatoporoid Coral Rudstone 7.7 0.1 16.8 N/A 0.01 N/A 0.1 0.01 6.6 26Coral Rudstone 4.5 0.5 9.9 2.2 0.03 82 1.0 0.01 147 41

DOLOSTONESSTRACHAN BUILDUP INTERIORSkeletal Wackestones 8.4 0.5 26 7.0 0.1 370 1.0 0.12 19 143Skeletal Packstones/Grainstones 7.1 4.8 12.6 15.0 1.2 396 5.1 0.43 367 13

RICINUS WEST MARGINTabular Stromatoporoid Boundstone 6.3 1.6 12.4 16.7 0.2 384 3.7 0.01 129 195Skeletal Wackestones 7.1 2.0 15.1 15.8 0.9 725 2.5 0.01 65 134Stromatoporoid Floatstone 7.7 2.0 15.3 27.9 1.9 501 2.4 0.2 38 89Coral Rudstone 6.6 3.0 12.4 18.5 2.3 262 2.0 0.2 28.4 30Breccias 6.5 3.3 12.0 21.1 1.7 525 1.5 0.2 9.6 19

RICINUS WEST INTERIOR (UPPER PORTION)Microbial Laminites 6.2 0.9 13.8 15.6 0.7 900 2.8 0.01 96 154Skeletal Wackestones 7.9 1.5 20.7 11.1 0.1 947 1.5 0.01 221 184Skeletal Packstones/Grainstones 6.1 1.3 15.0 9.5 0.06 753 1.9 0.01 142 313

RICINUS WEST INTERIOR (LOWER PORTION)Stromatoporoid Floatstone 6.0 0.6 15.5 12.7 0.1 764 1.7 0.01 100 228Coral Rudstone 5.6 1.0 15.2 11.9 1.6 418 1.5 0.01 41 70

*Ar = arithmetic mean; Geo = geometric mean; Kh = horizontial permeability; Kv = vertical permeability.

Table 6. Porosity and Permeability Ranges in Three Basic Pore Types, Ricinus West Reservoir.*

Thick- DistributionMain Porosity Porosity (%) Kh (md) Kv (md) ness Within Types Pore Types Ar σ n Min Max Geo Min Max Geo Min Max (m) Buildup

Amphipora Moldic Amphipora-like UpperMolds Intercrystalline 6.3 3.52 751 0.9 20.7 12.7** 0.06 2000 1.2** 0.01 861 131 Buildup

Fenestral-like Interior

Stromotoporoid Moldic pores Vugs (Stromatoporoid-like) Lower

Intercrystalline 6.1 3.00 294 0.6 19.1 13.1** 0.1 1400 1.8** 0.01 834 61 BuildupMoldic Interior(Thamnopora-like)

Large Vugs and Irregular vugs BuildupFractures Breccias 6.9 2.50 448 1.6 15.3 19.4** 0.2 2100 2.8** 0.01 141 259 Margin

Intercrystalline

*Ar = arithmetic mean, Geo = geometric mean, Kh = horizontal permeability, Kv = vertical permeability.**n of the geometric mean not equivalent to n of porosity average (values above max were dropped).

Page 293: Reservoir Quality Prediction in Sand and Carbonates

significantly higher than associated matrix based onminipermeameter measurements (N.C. Wardlaw, 1992,personal communication). Interconnection betweenpores is described as fracture-connected, as intercon-nected vugs/molds, and as matrix intercrystallineporosity (McNamara and Wardlaw, 1991). Lucia (1995)suggests that separate-vug and touching-vug porosityshould be distinguished. Separate vugs are typicallyfabric selective and include intrafossil, moldic, intra-granular, and shelter porosity. Lucia classifies touchingvugs as being typically nonfabric selective and includescavernous, breccia, fracture, solution-enlarged fracture,intercrystalline, and fenestral porosity. However, in thecase of these Upper Devonian buildups, most of thetouching vugs are solution-enlarged molds (e.g.,Amphipora) and are therefore fabric selective. IndividualAmphipora appear to overlap and connect with eachother in the horizontal direction. Specific pore types ofthe study area are discussed below.

Pore Types in Limestones

Limestones are present in pool D3B of the Strachanreservoir (wells 5-11, 14-2, 15-4) (Figure 2; Table 3).The facies distribution across pool D3B is illustrated incross section AA’ (Figure 3). Porosity and permeabilityplots are included with facies distribution to show therelationship between these reservoir parameters andthe depositional facies. Skeletal wackestone facieshave the highest porosity values (Table 5). Average Khand Kv of limestones are very low overall (0.8 and 0.4md, respectively). Pool D3B forms a poor reservoir.

The most common pore types in these limestonesare, in decreasing abundance, intraskeletal (50%–100%),subvertical fractures (5%–15%), and vugs (5%).Intraskeletal pores have spherical shapes in corals (Fig-ure 9A), tabular shapes in stromatoporoid fragments(Figure 9B), and are always lined with reservoir bitu-men (a descriptive term for bitumen that lines and fillspore spaces to distinguish it from bitumen in sourcebeds) (Rogers et al., 1974; Lomando, 1992). (In theremainder of the chapter, bitumen refers to reservoirbitumen.) Some interconnection is provided by theintraskeletal framework. Vugs are irregular in shape,small, and generally isolated. Minor subvertical frac-tures, with irregular surfaces that suggest dissolution,are filled with dolomite cement (C2, saddle dolomite;Figure 8), bitumen, and late-stage calcite (Figure 9C, D).

Pore Types in Partly Dolomitized Limestones

Partly dolomitized limestones are present in poolD3A of the Strachan reservoir (wells 12-31 and 10-31)(Figure 2; Table 3). Facies distribution and relation toporosity and permeability are illustrated in cross sec-tion BB’ (Figure 3). With 50 to 75 vol. % dolomite, alllime matrix is replaced, resulting in fine crystalline(30–60 µm) matrices with variable porosity rangingfrom porous intercrystalline to dense. Most skeletalfragments and marine cements remain as calcite (Fig-ure 6C). In partly dolomitized limestones, averageporosity is higher than in limestones (Table 5), but per-meability is low (see section on diagenetic control onpore systems). Stromatoporoid-coral rudstone and

skeletal wackestone facies have the highest averageporosity values, 7.7% and 4.9%, respectively; averageKh and Kv are higher in dolomites than in limestones(Kh 7.0 vs. 0.8 md, and Kv 1.0 vs. 0.01, respectively).

Pore types (Table 3) include intraskeletal (40%–80%),intercrystalline (5%–20%), vugs (5%) and subverticalfractures (5%–15%). Intraskeletal pores are restricted tostromatoporoid fragments (Table 3). Intercrystallinepores are polyhedral in shape, ~60 µm in diameter.Small to medium, irregular vugs are common in thedolomitized matrix of all facies (Figure 6D, E).

Pore Types in Dolostones

Dolostones are present in pool D3A of the Strachanreef interior (wells 7-32, 11-27, 11-22) (Figure 3) andthroughout the Ricinus West buildup. The facies distri-bution, porosity, and permeability values across theRicinus West reservoir are illustrated in cross sectionsAA’ and BB’ (Figure 5). With >75 vol. % dolomite, mostof the lime matrix is replaced, resulting in a densedolomite mosaic (Figures 6, 7, and 9D). The size of thedolomite crystals ranges from 60 to 250 µm, with somemuddy carbonates as fine as 30 µm. Most skeletalgrains have been dissolved, forming slightly enlargedmolds and irregular vugs, indicating a genetic relation-ship between the amount of dolomite and the moldicand vuggy pores (Figures 6, 9D–G; Table 4). Thesepores likely result from dissolution of calcite during orafter replacement dolomitization. Porosity (≤8.4% aver-age) and permeability are greater than in limestonesand partly dolomitized limestones (Table 5).

These dolostones contain varying amounts of vugs(40%–100%), molds (10%–60%), fenestral-like (10%–30%), intercrystalline (5%–15%), fractures and breccias(5%–10%), and minor (<1%) solution-enlargedintraskeletal pores (Figure 7; Table 4). Vugs show agradation from solution-enlarged molds to very large(3–7 cm), irregular pores that show little indication oftheir precursor fabric (Figure 9G). The most commonmoldic pores are those produced by selective dissolu-tion of tubular Amphipora fragments (e.g., well 7-32,4110.6 m) (Figures 6F, 7A, and 9F) in the upper part ofthe reef interior, referred to as Amphipora-like molds. Inthe lower part of the buildup interior, molds are com-mon after dissolution of domal (spherical) stromato-poroids and tubuar Thamnopora? fragments (Figure 9E),stromatoporoid-like and Thamnopora-like, respectively.Molds of Amphipora and Thamnopora fragments are dis-tinguished on the basis of their size, association withother pore types, and location within the buildup.Smaller pores that are tabular and aligned parallel tolaminations (fenestral-like) are considered molds relatedto microbial laminite facies (Figures 7E, F, and 10D).However, some dolomites have a nonporous, interlock-ing, crystal mosaic (Figure 10B).

Intercrystalline pores in fine-crystalline replacementdolomite are rare and very small (10 µm; e.g., RicinusWest, well 10-33, 4633.8 m). In contrast, intercrystallinepores in the more abundant, coarsely crystalline replace-ment dolomite are larger (250 µm; e.g., Ricinus West,well 10-33, 4535 m; well 15-23, 4488 m, Figure 10C, D).Thus, these dolostones show a positive relationshipbetween pore size and crystal size, as has been reported

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 283

Page 294: Reservoir Quality Prediction in Sand and Carbonates

284 Mountjoy and Marquez

elsewhere for other dolomites (Lucia, 1983, 1995; Cho-quette et al., 1992).

Fractures are commonly oriented in a subverticaldirection and are partly to completely filled bydolomite cement C1 and bitumen or late calcite cement(Figures 7D, 10E). Fracture width varies 1–5 mm, andlength ranges from 10 to 60 mm. Fractures are muchmore abundant in the buildup margins than in thebuildup interior (compare Figures 11, 12, Ricinus wells10-33 and well 7-13, respectively). Permeability is verysensitive to small changes in fracture porosity, whichin turn is largely dependent on fracture width (Lucia,1995). Dissolution along the fractures is indicated byassociated solution-enlarged intercrystalline andvuggy porosity. Brecciated intervals occur in thin, iso-lated zones to zones several meters thick in the RicinusWest buildup margin (e.g., well 7-13; Figure 10F). Brec-cias (Figure 10F) are characterized by centimeter-sizedolomite clasts with crackle, mosaic, and rubble tex-tures (Beales and Hardy, 1980) that are likely to be theresult of solution collapse (Amthor et al., 1993; Mar-quez, 1994). Matrices in the moldic breccias have inter-crystalline porosity (Figure 10F). Similar pore types arerecognized in dolomitized buildups (e.g., Westerose,Homeglen-Rimbey) along the Rimbey-Meadowbrookreef trend (Table 7) (McNamara and Wardlaw, 1991;Drivet, 1993). Bitumen-filled hairline microfracturesand other voids are relatively abundant in the Strachanbuildup. These microfractures are very late stagebecause they crosscut all diagenetic phases and extendsubhorizontally, radially, and randomly away fromvugs, molds, and fractures (Marquez and Mountjoy,1996). Pore space occluded by bitumen is discussedunder diagenetic controls and summarized in Table 8.

RESERVOIR CHARACTER IN DOLOSTONES

To evaluate the reservoir characteristics of dolostonebuildups in the deeper part of the Alberta basin, thedistribution and orientation of the different pore typesin the completely dolomitized Ricinus West buildupwere compared in three different parts of this buildupbecause of the excellent core available. The pore typeswere compared in three different parts of this buildupto determine the effect of facies on reservoir character.A number of different lithofacies (Figures 13, 14) occurin (1) the upper buildup interior (Upper Leduc Forma-tion), (2) the lower buildup interior (Middle Leduc),and (3) the buildup margin (Figure 13).

Upper Buildup Interior: Moldic and Intercrystalline Porosity

Three basic pore types constitute this part of thebuildup interior (in decreasing abundance) (Figure11): (1) moldic (Amphipora-like), (2) intercrystalline,and (3) fenestral-like pores (Figure 7). Moldic(Amphipora-like) pores are tubular, 10 mm to >3 cm(Table 4), and randomly oriented in a dense matrix.Effective communication is provided by intercon-nected molds. Intercrystalline pores are small, polyhe-dral in shape, well connected, and commonly lined

with a thin (1- to 4-µm) coating of bitumen. Fenestral-like pores are small, tabular in shape, and orientedparallel to laminations. These pores are interconnectedthrough intercrystalline pores in the matrix. Totalporosity ranges 0.9% to 20.7% (mean: 6.3%); Kh varies0.06 to 2000 md (geometric mean: 12.7 md); Kv ranges0.01 to 861 md (geometric mean: 1.2 md) (Table 6).

Lower Buildup Interior: Poorly Connected MoldsWith Some Intercrystalline Porosity

Three basic pore types (Figure 14) also form thereservoir in the lower buildup interior: (1) moldic (stro-matoporoid-like), (2) intercrystalline, and (3) moldic(Thamnopora-like). The lower buildup differs from theupper buildup interior because it contains larger andmore abundant molds of skeletal fragments. Moldic(stromatoporoid-like) pores are very large, spherical,and randomly oriented in a dense matrix. These poresare interpreted to represent the dissolution of domaland bulbous stromatoporoids. However, these vugs donot change the permeability significantly, becauseintercrystalline pores are similar to those of the upperbuildup interior. Additionally, >3-cm randomly ori-ented tubular pores, probably dissolved corals(Thamnopora-like), are common in a dense matrix. Totalporosity varies from 0.6% to 19.1% (mean: 6.1%); Khranges 0.1 to 1400 md (geometric mean: 13.1 md); Kvfrom 0.01 to 843 md (geometric mean: 1.8 md) (Table 6).

Buildup Margin: Random Vugs, Breccias, and Intercrystalline Porosity

Vugs, breccias, and intercrystalline porosity (Figure12) in decreasing abundance form the porosity typesalong the buildup margin (wells 7-13 and 11-27). Inter-connection is commonly provided by fractures andsome touching vugs. Total porosity ranges from 1.6%to 15.3% (mean: 6.9%); Kh ranges from 0.2 to 2100 md(geometric mean: 19.4 md); Kv from 0.01 to 141 md(geometric mean: 2.8 md) (Table 6).

SCALE, DISTRIBUTION, ANDVARIABILITY OF DIFFERENT

RESERVOIR TYPES

Large Scale (Tens of Meters to Kilometers)

The large-scale variations in porosity and perme-ability are shown in Ricinus West cross section, butoverall the porosity and permeability are remarkablyuniform (Figure 13). Although porosity and perme-ability are closely related to the depositional units andshow varations in Kh from 4.9 to 26.7 md (Figure 13B),data from each of these parts of the Ricinus West reser-voir indicate the lack of large-scale reservoir variabil-ity (Figure 13A; Table 6).

Medium Scale (Meters to Tens of Meters)

The differing arrangement of porosity and perme-ability in the three parts of the reservoir affects thepetrophysical properties of the reservoir. The verticaland lateral arrangement of the dominant pore types

Page 295: Reservoir Quality Prediction in Sand and Carbonates

vary considerably within these areas, and cause varia-tions at the medium scale in reservoir character. Thus,the Kh/Kv ratio is effectively much larger on this scalethan it is on the core-measurement scale. Within theRicinus West reservoir, several high- to medium-permeability intervals are separated by lower perme-ability zones (Figures 11–14; Marquez, 1994). In the

upper buildup interior, laterally continuous, high Khzones are associated with intercrystalline, moldic, and fenestral pores, whereas in the lower buildup inte-rior, Kh zones are related to Thamnopora-like and stro-matoporoid pores that are laterally discontinuous. In thebuildup margin, Kv zones are common and are apparentlyrelated to subvertical fractures. Vertical permeability

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 285

Figure 9. Core and thin-section photographs from Strachan and Ricinus buildups. Scales for cores are in centime-ters. (A) Intraskeletal pores in calcite coral fragment. Strachan D3B 15-2-38-10W5, 4113.4 m. (B) Dolomitized tabu-lar stromatoporoid boundstone facies with solution-enlarged intraskeletal pores (arrows) lined with reservoirbitumen. Some vugs filled with calcitized anhydrite. Ricinus West 7-13-36-10W5, 4621.3 m. (C) Porous, mediumcrystalline (62–250 µm), planar, euhedral replacement dolomite. Intercrystalline pores are lined with reservoirbitumen (black). Strachan D3A 10-31-37-9W5, 4287.8 m. Scale bar is 50 µm. (D) Nonporous replacement dolomitewith irregular vugs (V). Stylolites (arrows) connect vugs, since they are filled with reservoir bitumen (black). Latecalcite cement (Ca) partly fills vug. Ricinus West 10-33-36-10W5, 4601.5 m. Scale bar is 200 µm. (E) Thamnopora?molds (C) in a tight dolomite, coral rudstone facies. Some molds are solution enlarged. A few microfractures.Ricinus West, 7-13-36-10W5; 4514 m. (F) Tubular moldic pores probably after Thamnopora fragments in a tightmatrix of replacement dolomite. Some small fractures (arrow) connect molds. (G) Large irregular vugs in replace-ment dolomite. Strachan D3A 12-31-37-9W5, 4295 m.

Page 296: Reservoir Quality Prediction in Sand and Carbonates

286 Mountjoy and Marquez

profiles (Figures 11, 12) indicate that the Ricinus Westreservoir can be further subdivided into thinner faciesand porosity-permeability slices as indicated in Figure13B. In the reef interior, a lower sequence with later-ally discontinuous permeable zones is overlain by astacked sequence of thin, laterally continuous perme-able zones with poor vertical interconnection. In thereef margin, thick, laterally discontinuous zones arecharacterized by higher Kv (Figure 12), apparentlyrelated to subvertical fractures. These data have impli-cations not only for the understanding of reservoirheterogeneity and character, but also for the predic-tion of reservoir quality.

Small Scale (Meters to Millimeters)

At this small scale, porosity and permeability arecontrolled by the facies, individual pore types, anddifferences in diagenesis. In intercrystalline pores

(Figure 11C) of the upper buildup interior, the pres-ence of 10%–15% bitumen (percentage in terms ofbulk volume) (Table 8) sufficiently lines most porethroats and thus greatly reduces the porosity and per-meability. In stromatoporoid-like and Thamnopora-like molds in the lower buildup interior (Figure 14),bitumen is less abundant (~2%), and permeabilityvariations are controlled by the presence of minordolomite, anhydrite, and calcite cements. Bitumen-filled hairline microfractures are abundant in vugs inthe upper part of the reef margin (above a paleo-oil-water contact); therefore, porosity and permeabilityare mainly controlled by the presence of subverticalfractures and vugs (Marquez and Mountjoy, 1996). Inthe lower part of the reef margin, touching vugs (withminor bitumen and cements) and some fractures pro-vide Kv. The effects of bitumen, cementation, and dis-solution are discussed below.

Figure 10. Core and thin-section photographs from dolomitized Ricinus buildup. Scales for cores are in cen-timeters. (A) Nonporous replacement dolomite with solution-enlarged pores (V), probably after coral frag-ments. Ricinus West 15-23-36-10W5, 4592.9 m. Scale bar 250 µm. (B) Nonporous mosaic of subhedral, mediumcrystalline, planar replacement dolomite. Crimson 10-24-37-9W5, 4032.3 m. Scale bar 100 µm. (C) Polyhedralintercrystalline pores (dark) in replacement dolomite lined by reservoir bitumen (black). Plane light. RicinusWest 15-23-36-10W5, 4488.4 m. Scale bar 50 µm. (D) Microbial laminite dolostone with irregular vugs in a tightmatrix. Reservoir bitumen (black) lines vugs. Plane light. Ricinus West 11-27-36-10W5, 4490.8 m. Scale bar250 µm. (E) Intercrystalline porosity in replacement dolomite adjacent to subvertical fracture partly dolomitecement. Ricinus West 10-33-36-10W5, 4621.6 m. (F) Dolomite breccia and fracture porosity within intercrys-talline porosity. Pores are lined with bitumen. Ricinus West 7-13-36-10W5, 4393.5 m.

Page 297: Reservoir Quality Prediction in Sand and Carbonates

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 287

4546

4547

4548

4549

4550

4551

4552

4553

4554

F

4555

4556

4557

4558

4559

4560

4561

4562

4563

4537

4538

4539

4540

4541

4536

4543

4544

4545

4542

A) Depositional Facies: SKELETAL PACKSTONES GRAINSTONES Dominant Pore Type: Amphipora-like Molds

B) Depositional Facies: LAMINITE

Dominant Pore Type: Fenestral-like

C) Depositional facies: SKELETAL WACKESTONE

Dominant Pore Type: Intercrystalline

Irregular vugsStylolites

F Fractures

Fenestral-likelaminations

Fenestral-likepores

Amphipora-likepores

Horizontal permeability

Depth (m) Porosity (%) Kh Kv Permeability (ms) (md) Zone

10

High Low

No DataMedium

Intercrystalline pores

Vertical permeability

Figure 11. Examples of low-, medium-, and high-permeability zones in a 27-m-thick sequence in upperbuildup interior, Ricinus West buildup interior (well 10-33; see Figure 5 for location of this interval). Thin,high-permeability zones are separated by thicker intervals of lower permeability. (A) Moldic (Amphipora-like)pores in skeletal packstone and grainstone facies. (B) Fenestral-like pores in microbial laminite facies. (C)Intercrystalline pores in skeletal wackestone facies. Kh = horizontal permeability, Kv = vertical permeability.

Page 298: Reservoir Quality Prediction in Sand and Carbonates

288 Mountjoy and Marquez

F

Dominant Pore Type: Vugs and/or BrecciasDepositional Facies: TABULAR STROMATOPOROID BOUNDSTONES

F

4 4 1 5

4 4 2 0

4 4 2 5

4 4 3 0

4 4 3 5

4 4 4 0

Depth (m) Porosity (%) Kh (md) Kv (md) Permeability zones 0 6 12 0 10 200 500 0 30

Shaly interval

Vertical permeability

Fractures

Horizontal permeability

High

Medium

No Data

Low

F

Stylolites

Irregular vugs

Hairline microfracturesextending from vugs

Figure 12. Examples of high-, medium-, and low-permeability zones in buildup margin, Ricinus West (well7-13; see Figure 5 for location of this interval). In the lower 11 m, vertical permeability (Kv) is provided byconnecting vugs and some fractures. In the upper 15 m, vertical permeability is dominated by partly opensubvertical fractures. Kh = horizontal permeability.

Page 299: Reservoir Quality Prediction in Sand and Carbonates

DEPOSITIONAL CONTROLS ON PORE SYSTEMS

Reservoir interconnection is the most importantparameter in reservoir simulation studies (Van deGraff and Ealey, 1989), and is related to the presenceand distribution of internal vertical and horizontalchanges in permeability (Ghosh and Friedman, 1989).These changes in permeability are plotted in verticalprofiles for the three parts of the West Ricinus reser-voir (Figures 11, 12, and 14). Horizontal permeabilitydata have a wide range of values and therefore havearbitrarily been grouped into high (>200 md), medium(10–200 md), and low (<10 md). Vertical permeabili-ties >10 md are considered high.

In the upper buildup interior, the vertical successionof Amphipora-like molds and fenestral-like and intercrys-talline pores resemble the shallowing-upward deposi-tional parasequences developed in the upper reefinterior. These intervals range in thickness from 8 to 27 m(see well 10-33 in Figure 5). The vertical permeabilityprofiles in Figure 11 show that in a 27-m-thick sequencecontaining Amphipora-like (Figure 11A), fenestral-like(Figure 11B), and intercrystalline pores (Figures 5, 8, and11C), thin (1–2 m), high Kh zones are separated bythicker intervals (2–5 m) of lower permeability. Thisindicates that within each sequence, vertical intercon-nection is poor, and horizontal interconnection is pro-vided by relatively thin permeable zones that wouldform fluid flow conduits. Slightly thicker permeabilityzones occur where intercrystalline porosity dominates.Based on cross sections AA’ and BB’ (Figure 5), deposi-tional models for limestone buildups (McGillivray andMountjoy, 1975) and outcrop studies of equivalent strata(McLean and Mountjoy, 1993a, b, 1994), these deposi-tional sequences are extensive and laterally continuous.The upper buildup interior is characterized by six 8- to27-m-thick intervals of different pore types that closelyfollow the depositional sequences, with poor verticalinterconnection and thin, laterally continuous, highlypermeable horizontal zones. Such thin, porous, and per-meable laterally continuous strata would form conduitsor flow units along which higher flows would occur.

In the lower buildup interior, vertical permeabilityis slightly higher (Figure 14). Horizontal permeabilityzones are slightly thicker where Thamnopora-likemolds dominate (Figures 9F, 14B), with connectivity

provided by touching molds. From previous studies(McGillivray and Mountjoy, 1975; McLean, 1992),facies associated with stromatoporoid and coral moldsare known to be laterally discontinuous. This part ofthe buildup interior consists of thin, laterally discon-tinuous, porous, and permeable zones.

In the buildup margin, vertical permeability is com-mon and associated with subvertical fractures (Figure 12,4423 m) and connecting vugs (Figure 12, 4439 m). Hori-zontal permeability zones are thin (<1 m) and laterallydiscontinuous. Brecciated intervals that contribute toreservoir interconnection (see cross section BB’, Figure 5)are common in the reef margin. Thus, the buildup mar-gin is characterized by thick intervals where Kv predom-inates. This is similar for the Homeglen-Rimbey buildupmargin (Drivet, 1993; Drivet and Mountjoy, 1997).

In summary, the buildup interior reservoir changesupward from laterally discontinuous Thamnoporamoldic pores in a dolomite matrix with intercrystallineporosity to a similar matrix dolomite with Amphiporaand fenestral-like pores in 1- to 2-m-thick units withhigh Kh separated by thicker intervals of dolomite withlower permeability. The buildup margin has better Kvdue to subvertical fractures that connect the vugs andbreccia zones. As noted above, the measured averagematrix porosities and permeabilities in all three regionsof the reservoir are similar and appear to be mainly aresult of the dominant control on reservoir character byintercrystalline porosity in the matrix dolomites. How-ever, considerable variation must occur in the porosityin different parts of the reservoir as a result of varia-tions in moldic, vuggy, and breccia porosity.

DIAGENETIC CONTROLS ON PORE SYSTEMS

Dolomitization, cementation, dissolution, pressuresolution, and bitumen plugging affect the pore sys-tems and the relationship between porosity and per-meability differently.

Dolomitization and Porosity

The degree of dolomitization in skeletal wackestonesfacies was investigated to examine the effects of dolomiti-zation on porosity and other diagenetic effects (Figures

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 289

Table 7. Porosity and Permeability in Limestone Buildups Along the Rimbey-Meadowbrook Reef Trend.

Golden Spike* Strachan (Pool D3B)(<2000 m) (>4000 m)

Depositional Facies Porosity Porosity PermeabilityReef Margin Mean (%) Kh (md) Mean (%) Kh (md) Kv (md)

Coral Rudstones 4 1 3 0.3 0.01Tabular StromatoporoidBoundstones 5.5 10 2.3 0.1 0.02

StomatoporoidCoral Rudstones 10 100 2.6 0.1 0.01

*From Walls and Burrowes (1985).Kh = horizontal permeability, Kv = vertical permeability.

Page 300: Reservoir Quality Prediction in Sand and Carbonates

290 Mountjoy and Marquez

Tab

le 8

. Eff

ect o

f R

eser

voir

Bit

um

en o

n P

oros

ity

and

Per

mea

bil

ity

in D

olom

itiz

ed S

kel

etal

Wac

kes

ton

e Fa

cies

, Ric

inu

s W

est R

eser

voir

.

Page 301: Reservoir Quality Prediction in Sand and Carbonates

15–18). Skeletal wackestones were chosen because theyhave the most homogeneous depositional texture and arethe most widespread facies throughout the buildups. Inaddition, except for the presence of ~10% bitumen, otherlate-stage diagenetic products such as dissolution vugs

and late calcite are minor components in these wacke-stones. Skeletal wackestones with <50 vol. % dolomitehave porosities ranging from 1% to 5% (mean: 2.4%; Fig-ure 16A, B). Completely dolomitized skeletal wacke-stones range in porosity from 0.5% to 14% (mean: 5.2%;

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 291

0 1 2

Km

!00

0

m

A)

B)

0 1 2

Km

100

0

m

?

UPPER BUILDUP INTERIOR=6.3% Kv=1.2md

Kh=12.7 md

LOWER BUILDUP INTERIOR =6.1% Kv=1.8md Kh=13.1 md

?

?

Kh = 8.7 md Kv = 1.2 md6 %

Kh = 15.1 md Kv = 1.0 md7 %

6 % Kh = 9.2 md Kv = 0.8 md

7 %

5 % Kh = 4.9 md Kv = 0.6 md

Kh = 16.3 md Kv = 0.4 md

7 % Kh = 26.7 md Kv = 03 md

7 % Kh = 15.1 md Kv = 1.7 md

6 % Kh = 16.4 md Kv = 2.4 md

6 % Kh = 21.5 md Kv = 2.7 md

4 % Kh = 8.7 md Kv = 1.4 md

10-3311-27

7-13

6-257-26

BUILDUP INTERIOR WELL

BUILDUP MARGIN WELL

?

BU

ILD

UP

MA

RG

INB

UIL

DU

P M

AR

GIN

BU

ILD

UP

MA

RG

IN

6 % Kh = 6.5 md Kv = 0.9 md

BU

ILD

UP

MA

RG

IN

COOKING LAKE Fm.

= 6.92%Kv = 7.44 mdKh = 49.56 md

= 6.9% Kv = 2.8 mdKh = 19.4 md

IRETON Fm.

IRETON Fm.

DUVERNAYFm.

DUVERNAYFm.

UPPER LEDUC

MIDDLE LEDUCIRETON Fm.

LOWER BUILDUPINTERIOR

IRETON Fm.

IRETON Fm.

DUVERNAY Fm.

COOKING LAKE Fm.

UP

PE

R B

UIL

DU

PIN

TE

RIO

R

DUVERNAY Fm.

IRETON Fm.

Figure 13. Subdivision of the Ricinus West reservoir in three large-scale regions; upper and lowerbuildup interior and buildup margin. (A) Average porosity (6%) is similar throughout the entirereservoir. Horizontal and vertical permeabilities are slightly higher in the buildup margin. (B)Medium-scale porosity and permeability subdivisions within the buildup interior. The upperbuildup interior is subdivided into six zones, 8 to 27 m thick. Horizontal permeability in each zone ishigh, but vertical permeabilities are low. See Figure 11 for details of pore types and depositionalfacies. The lower buildup interior has laterally discontinuous porosity and permeability. The thickbuildup margin has greater vertical permeability.

Page 302: Reservoir Quality Prediction in Sand and Carbonates

292 Mountjoy and Marquez

A)

B)

Dep

ositi

onal

Fac

ies:

ST

RO

MA

TO

POR

OID

FL

OA

TST

ON

ES-

R

UD

STO

NE

S

Dom

inan

t Por

e T

ype:

Str

omat

opor

oid-

like

mol

dsD

omin

ant P

ore

Typ

e: T

ham

nopo

ra-l

ike

mol

dsD

epos

ition

al F

acie

s: C

OR

AL

RU

DST

ON

ES

LO

WE

R B

UIL

DU

P I

NT

ER

IOR

: W

ell

10-3

3 R

ICIN

US

Hor

izon

tal p

erm

eabi

lity

Hig

h

Med

ium

Low

No

Dat

a

Styl

olite

s

Irre

gula

r vu

gsA

nhyd

rite

-fill

edir

regu

lar

vugs

Inte

rcry

stal

line

pore

s

Ver

tical

per

mea

bilit

y

Shal

y in

terv

al

Stro

mat

opor

oid-

like

pore

s

Tha

mno

pora

-lik

epo

res

Figu

re 1

4. E

xam

ple

s of

low

-, m

ediu

m-,

and

hig

h-p

erm

eab

ilit

y zo

nes

in lo

wer

bu

ild

up

inte

rior

, Ric

inu

s W

est b

uil

du

p (w

ell 1

0-33

; see

Figu

re 5

for

loca

tion

of

this

inte

rval

). H

igh

er p

erm

eab

ilit

y zo

nes

are

thic

ker

in th

e co

ral r

ud

ston

e fa

cies

wit

h a

bu

nd

ant T

ham

nopo

ra-l

ike

por

es. D

omin

ant p

ore

typ

es: (

A) s

trom

atop

oroi

d-l

ike

mol

ds

in s

trom

atop

oroi

d f

loat

ston

es/r

ud

ston

es,

(B) T

ham

nopo

ra-l

ike

mol

ds

in c

oral

ru

dst

one

faci

es. K

h=

hor

izon

tal p

erm

eab

ilit

y, K

v=

ver

tica

l per

mea

bil

ity.

Page 303: Reservoir Quality Prediction in Sand and Carbonates

Figure 16C) and have considerably higher Kh (≤100 md;Figure 16F). Thus, porosity increases from <5% in par-tially dolomitized limestones, on average, to ≤10% to 14%in completely dolomitized facies (Figure 16). Completelydolomitized skeletal wackestones in the reef interior ofpool D3A (Figure 17A–C) have the highest porosities inthe entire buildup, ranging from 3.4% to 25% (mean:±9.8%) and much higher permeabilities (Figure 17D–E)than in partially dolomitized limestones. This is clearlyshown by the log plot of percent dolomite and porosity[Figure 18, calculated from sonic and density logs usingthe Chaveroo method (Bateman, 1985) and assumingfrom core logging that the lithology was dominantlylimestone and dolomite]. There is a satisfactory matchwith estimated amounts of dolomite and measuredporosities from Strachan well 10-31, D3A pool (Figure 18).In the partially dolomitized intervals where limestone is≥60%, the porosity is very low. Not all dolomites areporous, as shown by the low porosities in the completelydolomitized sections, especially in the interval4179–4186 m. Porosity in this case is a combination ofintercrystalline and vuggy and is best developed in stratawith >50%–60% dolomite, with the highest porositiesoccurring in completely dolomitized rocks. Theseincreases in porosity appear to result mainly from disso-lution (D1) of calcitic skeletal fragments and grains thatwere either associated with dolomitization or occurredlater. Probably the development of intercrystalline poros-ity in the matrix also increased the porosity. The increasedabundance of molds, solution-enlarged molds, and irreg-ular vugs and porosity may not necessarily be related todolomitization itself, but may in part be due to later solu-tion events (Figure 8).

Dolomitization and Permeability

Dolomitization greatly affects permeability, but onlywhen the rocks are 100% dolomite, which have highestabsolute and average horizontal permeability (Figures15–17D–F; geometric mean Kh: 27.9 md) and verticalpermeability values [geometric mean Kv: 5.1 md (Table5)]. As the percentage of dolomite increases, no signfi-cant trends occur in the permeabilities in these samplesfrom the Strachan reservoir (Figures 6A–C; 15–17), butthere is a large scatter in the data. Skeletal wackestoneswith <50 vol. % dolomite content tend to have very lowhorizontal permeabilities (0.01–0.4 md; Figure 16D, E).Completely dolomitized skeletal wackestones in the reefmargin have greater horizontal permeabilities, rangingfrom 0.01 to 200 md (Figure 16F). Completely dolomi-tized skeletal wackestones in the Strachan buildup inte-rior have high permeabilities from 0.01 to 450 md(Figure 17 D–F). Similar porosity and permeabilitiesoccur in the completely dolomitized Ricinus reef margin7-13 well (Figures 19, 20; compare Figure 17A with 19C).The crossplots of porosity with permeability from Rici-nus 7-13 show considerable variation. In general, there isan overall increase in permeability with increasingporosity, which may be related to increasing crystal (orparticle) size, as suggested by Lucia (1995). In Ricinus 7-13, below 4430 m there are several intervals with high Kv(Figure 12) that appear to be related to fracturing.

Moldic vugs can increase porosity, but if not connectedwill not contribute significantly to increased perme-ability.

Comparing the porosity-permeability plots with thethree permeability classes of Lucia (1995) based on parti-cle size indicates that the Devonian limestones anddolomites of the Strachan and Ricinus reservoirs fallwithin or above the >100-µm particle size (Figures 21,22) [Lucia’s (1995) rock-fabric/petrophysical class 1].Both the dolomitized limestones (Strachan 10-31) andthe 100% dolomitized intervals exhibit permeability val-ues higher than expected. The limestones have perme-abilities ≤10 md (Figure 21), whereas the dolomites havesome permeability values slightly >1000 md (Figure 22).The reasons for permeabilities higher than expected isprobably due to small and large vugs being well inter-connected with each other via matrix intercrystallineporosity, and perhaps some fracturing. Interestingly,there is little difference in the porosity-permeabilityplots in the Ricinus buildup (Figure 22) between thebuildup margin (7-13) and the buildup interior wells(10-33), even though there are more fractures in thebuildup margin wells (cores with fractures tend tobreak, and it is difficult to measure their porosity andpermeability). Thus, these dolomites and partiallydolomitized strata with touching or interconnected vugsform excellent reservoirs with good horizontal perme-abilities. Permeability appears to increase with dolomitecrystal size, but this relationship has not been studiedsystematically in these reservoirs. Dolomitization of pre-viously mud-dominated rocks tends to increase perme-ability and to some extent porosity.

Lucia (1995) considers that all touching vugs (frac-tures, breccia, caverns, fenestral) are diagenetic in originand are unrelated to primary depositional features.Touching vugs in the Devonian strata of westernCanada are abundant in all dolomitized reservoirs. Par-tial to complete dissolution of Amphipora and stromato-poroids formed a series of molds and solution-enlargedmolds that make up touching or connected vugs wherethese fossils were abundant, especially along thebuildup margins and in the adjacent lagoons. Thesevugs are interconnected where they touch adjacent vugsand via the medium to coarse intercrystalline porosity inthe surrounding medium to coarsely crystalline matrixdolomites, forming a laterally extensive three-dimen-sional conduit for fluid flow. Hence, a moldic classrelated to primary textures needs to be added to Lucia’s(1995, his figure 17) touching-vug category.

It is critical to determine whether such moldic vugsare physically separate or touching (connected). If sepa-rate, these vugs increase total porosity but do not signif-icantly increase permeability. However, if they aretouching, they increase permeability (Figure 22), andmost of the porosity would be effective. Unlike thetouching vugs classified by Lucia (1995) as being whollydiagenetic and unrelated to depositional fabrics, theseDevonian moldic vugs are related to the depositionalfabrics. Such fabrics are more readily characterized, andthe resulting porosity and permeability are more easilymapped and extrapolated in these Devonian strata; inshort, porosity and permeability are more predictable.

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 293

Page 304: Reservoir Quality Prediction in Sand and Carbonates

294 Mountjoy and Marquez

Thus, although all depositional facies have beenmodified by several diagenetic processes such as earlycementation, chemical compaction, and later dissolu-tion and cementation, replacement dolomitization andassociated dissolution exert the most pronouncedeffect on reservoir fabrics (Figures 16–20). This effect istwofold: it results in significantly higher porosities and,more importantly, significant increases in permeabili-ties, thus enhancing their capacity to act as flow units.

Effect of Cementation and Dissolution

In Ricinus West, cementation occurred during morethan one stage (Figure 8). The net effect of cementation isnot only to reduce the pore size but to reduce or com-pletely plug pore throats (Wardlaw, 1980), resulting in adrastic reduction of permeability. Anhydrite cementa-tion is volumetrically important in the lower part of thereservoir. At least three dissolution events modified theRicinus West reservoir rock. Dissolution D1 (Figure 8)affected calcite skeletal fragments during or after replace-ment dolomitization, resulting in molds, solution-enlarged molds, and vugs in the entire reservoir,apparently improving the connectivity and permeability(Figures 6E, F, 7H, and 9B). Dissolution D2 and minor

dissolution D3 of matrix dolomite resulted in vugs andbrecciated intervals that are more abundant in the reefmargin. The distribution of these cements and occur-rences and amounts of dissolution are difficult to predict.

Effects of Pressure Solution

The effects of stylolitization and other pressuresolution phenomena in Leduc carbonates have beendiscussed by Mossop (1972) and Amthor et al. (1994),among others. The equivalent limestone facies in theGolden Spike buildup (1600 m present depth) havemuch higher porosities and permeabilities than lime-stones in the Strachan buildup (Table 7). This dramaticdecrease of porosity and permeability is primarily theresult of pressure solution (Figure 7F, I) during anadditional 2 km of burial in the deeper basin, togetherwith pore filling by later cements and bitumen. Evenin the shallower part of the basin, pressure solutionhas caused considerable reduction in porosity in lime-stones and produced secondary effects such as theraised rim of the Redwater buildup (Mossop, 1972).

Effect of Bitumen on Porosity and Permeability

Bitumen has been reported as a common porosity-occluding phase in some reservoirs (McCaffery, 1977;Lomando, 1992). The precipitation of bitumen cancause wettability characteristics (the relationshipbetween hydrocarbon-rock contact angles) to changefrom water-wet to mixed or strongly oil-wet(Lomando, 1992). For example, McCaffery (1977)reported that pore-lining bitumen in the DevonianWindfall D-3 pool made the rock intermediately wetfor gas-water systems. Wettability strongly influencesreservoir behavior and ultimate hydrocarbon recovery(Wardlaw, 1992). Therefore, the type, amount, and dis-tribution of bitumen needs to be carefully evaluated.

In the Strachan and Ricinus West buildups, deposi-tional facies with different pore types have variableamounts of bitumen with homogeneous to heteroge-neous distribution (Figures 6D, F; 7C, D; 9B–D; and 10C,D). Skeletal wackestones have the highest percentages(15%) of bitumen and are characterized by abundantintercrystalline pores (Figures 9C, 10C), some smallvugs, and, locally, fractures. All pore types are lined topartly filled with bitumen as thin (1–4 µm) coats anddroplets (≤10 µm). Bitumen is most common in theupper part of reservoirs, mainly above the gas-watercontact. The prebitumen porosity (visual estimates) andpresent (core-measured) porosity for this facies areshown in Table 8. In well 10-33, skeletal wackestonesinterbedded with microbial laminites in the upper reefinterior of the Ricinus West reservoir have prebitumenporosity of 12.9% that has been reduced to 2.9% (a 70%reduction) after emplacement of an average bitumencontent of 10% (at 4601 m). A similar porosity reductionoccurs at 4561.5 m. In the lower reef interior, skeletalwackestones are interbedded with domal stromato-poroid facies and have a relative low bitumen content(2%; 4628.9 m), which reduced porosity by 22%. Similarreductions in porosity occur in well 11-27 (Table 8).Thus, precipitation of bitumen in the Ricinus Westreservoir reduces porosity consideraby.

10090807060504030201001

10

100

1000

Dolomite (%)

10090807060504030201000

5

10

15

20

Dolomite (%)

A)

B)

0

Figure 15. Core-measured porosity (A) and perme-ability (B) vs. dolomite content in all depositionalfacies of Strachan well 10-31 (pool D3A; see Figure 3for location and facies distribution).

Page 305: Reservoir Quality Prediction in Sand and Carbonates

The greatest significance of bitumen precipitation onreservoir character is reduction of permeability andincrease in reservoir heterogeneity. Comparison of corepermeability in a continuous, 1-m-thick interval ofskeletal wackestone facies (well 11-27, 4478 m; Table 8)illustrates the effect of bitumen plugging on permeabil-ity. With 10% bitumen, Kh averages 3.4 md and Kv aver-ages 0.4 md, compared to 45.8 md and 24.8 md,respectively, in dolostones with 1% bitumen. This

reduction in permeability is caused by thin coats ordroplets of bitumen that restrict or completely blockpore throats. This bitumen was probably related to thedeasphalting of the oil. This late-stage deep burial eventimparts a component of reservoir heterogeneity thatmostly is unrelated to depositional facies or prebitumendiagenesis. In addition, where overpressuring tookplace in an isolated or a sealed reservoir such as Stra-chan during the conversion of crude oil to gas beginning

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 295

0 2 4 6 8 10 12 14

Porosity (%)

Strachan Pool D3BWell: 14-2 4116-4119m

Dolomitization: 30%Mean: 2.8Std. deviation: 1.2n: 9

Stachan Pool D3B

A)

0

2

4

6

8

10

12

14

16

18

20

0

2

4

6

8

10

12

14

16

18

20

0 2 4 6 8 10 12 14

4162-4181 mWell: 10-31

25

StrachanPool D3A

C)

0 2 4 6 8 10 12 14

B)Strachan Pool D3AWell: 10-314115-4119 m

Dolomitization: 20%Mean: 2.1Std. deviation: 0.79n: 16

0

2

4

6

8

10

12

14

16

18

20

0

5

1 0

1 5

2 0

2 5

3 0

3 5F)

0 1 2 3 4 5 6 7 8 9 10

0 10 20 30 40 50 60 70 80 90 100

0

5

1 0

1 5

2 0

2 5

E)

0

5

1 0

1 5

2 0

2 5

0 2 3 4 5 6 7 8 9 10

D)

n

n

n

n

n

n

Porosity (%)

Porosity (%)

Dolomitization: 100% Mean: 5.2 Std. deviation: 206 n:72

1

Well 14-24116-4119 m depthDolomitization: 30%n=11

Strachan Pool D3AWell 10-314115-4119 m depthDolomitization: 20%n=11

Pool D3AWell 10-314162-4181 m Dolomitization: 100%n=68

Figure 16. Porosity and permeability histograms measured from skeletal wackestone facies Strachan buildupshowing variations in reef margin D3A and D3B pools: A–D, limestones (<50 vol. % dolomite content); E–F,dolostones (>75 vol. % dolomite content). See Figure 3 for well locations.

Page 306: Reservoir Quality Prediction in Sand and Carbonates

296 Mountjoy and Marquez

at temperatures above ~120°C, microfractures wereformed and bitumen was forced into most pore throatsand fractures (Marquez and Mountjoy, 1996).

Porosity Evolution in the Strachan and Ricinus West Buildups

Evolution of the pore systems is interpreted tohave occurred in four major stages (Figure 23) that

represent important changes in porosity and perme-ability of the reservoirs.

Stage 1: Deposition (Facies-Controlled Original Porosity)

The paleogeographic setting originally controlleddepositional environments and, ultimately, the origi-nal porosity and permeability. Skeletal boundstonesand rudstones with packstone matrices and marine

Well: 11-27Strachan Pool D3A

3927-4023 mMean: 8.8

n=32Std. deviation: 2.7

Dolomitization: 100%

B)

Well: 11-223953-3977 m

StrachanMean: 10.9Std. deviation: 3.4n=10

Dolomitization: 100%Pool D3A

1 count at 18-20%

C)

A)

Well: 7-Strachan Pool D3A

4083-4112 mMean: 8.8

n=4Std. deviation: 4.3

0 2 4 6 8 10 12 14 16

0 2 4 6 8 10 12 14 16

0 2 4 6 8 10 12 14 16

Dolomitization:100%

1 count at 26-28%

Well: 7-32Strachan Pool D3A

4083-4116 mDolomitization: 100%n=43

4 counts above 250 md

D)

Strachan Pool D3AWell: 11-273927-4023 mDolomitization: 100%n=31

4 counts above 250 md

E)

20Well: 11-22

n=11

Strachan Pool D3A

3954-3961 mDolomitization: 100%

25

F)

0 50 100 150 200 225

32

20

18

16

14

12

10

8

6

4

2

0

20

18

16

14

12

10

8

6

4

2

0

20

18

16

14

12

10

8

6

4

2

0

Porosity (%)

Porosity (%)

Porosity (%)

n

n

n n

n

n

15

10

5

20

25

0 50 100 150 200 225

15

10

5

20

25

0 50 100 150 200 225

15

10

5

0

Kh (md)

Kh (md)

Kh (md)

0

0

Figure 17. Porosity and permeability histograms measured from 100% dolomitized skeletal wackestone faciesin wells 7-32, 11-27, and 11-22, interior of Strachan buildup pool D3A. See Figure 3 for well locations.

Page 307: Reservoir Quality Prediction in Sand and Carbonates

cements were deposited at the reef margins. Skeletalrudstones, grainstones, packstones, wackestones, andmudstones were deposited in the reef interiors. Initialporosities were probably high and differed among thedifferent facies. For example, modern packstones andgrainstones from Florida have porosities from 40% to67% and permeabilities from 1840 to 30,800 md,whereas wackestones and mudstones have higherporosities (68%) but lower permeabilities, 228 to 0.87 md(Enos and Sawatsky, 1981). Thus, depositional envi-ronments strongly control initial porosity and perme-ability. This primary porosity was reduced by near seafloor and early burial calcite cementation, especiallyalong the buildup margins (Figures 3, 8). During bur-ial, the sediments were subjected to mechanical andchemical compaction.

Stage 2: Replacement Dolomitization (Early Burial,Porosity Reduction)

During the early stages of replacement dolomitiza-tion (as observed in the Strachan buildup), isolateddolomite rhombs nucleated and grew in lime mud ofall depositional facies (stage 2A, Figure 23). Asdolomitization progressed, dolomite rhombs grew toform dolomite patches in the lime mud matrix (stage2B) of rocks in the D3A pool, and also began fillingsome porosity. Less dolomite formed in the D3B pool(column 1, Figure 23). In places, advanced matrix

dolomitization formed a self-supporting frameworkthat probably made these rocks more resistant tochemical compaction (stage 2C). Conversely, lime-stones were subjected to continuing pressure solutionand porosity reduction (column 1, Figure 23). Duringor following replacement dolomitization, porosity wasrearranged with dissolution (D1) of allochems, form-ing vugs and increasing porosity (stage 2D).

Stage 3: Cementation, and Stage 4: Dissolution andBitumen Emplacement

Either toward the end of fossil dissolution or after-ward, minor amounts of dolomite cement (C1) locallydecreased porosity. Dolomite cement (C2), anhydrite,and minor sulfides partly filled pores, mostly in thelower part of the reservoirs. The emplacement of bitu-men during the Late Cretaceous notably reducedporosity and, more importantly, permeability, in theupper part of the reservoir.

DISCUSSIONControls on Pore Types and Their Distribution

In carbonate reservoirs like the Leduc buildups,most depositional facies can be permeable or nonper-meable, depending on the type and extent of diagene-sis. Porosity and permeability variations in theStrachan and Ricinus West buildups are controlled by a

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 297

Figure 18. Partially dolomi-tized interval in Strachanwell 10-31, D3A pool, showing (on the right) therelative amounts of limestone (blue), dolomite(magenta), and porosity(red), derived from equations(Chaveroo method) usingsimultaneous equations ofsonic and density data (runin 1968) utilizing GEOGRAPHIXQLA2 log analysis software.On left, porosity and permeability curves; blue,measured core porosities andpermeabilities (ERCB data),and brown, log calculatedporosities and permeabilities.Porosity increases withincreasing dolomitization.Highest porosities occur onlyin completely dolomitizedstrata. In general, there is agood match between the log-calculated porosity andlithology, except that theamount of dolomitecalculated from logs is20%–40% higher than thatestimated from the cores.

Page 308: Reservoir Quality Prediction in Sand and Carbonates

298 Mountjoy and Marquez

series of complex interactions involving depositionalenvironments and multiple diagenetic events duringprogressive burial, thermal maturation, overpressuring,and thermochemical sulfate reduction. The upper reefinterior is characterized by porous, permeable, laterallycontinuous slices that range in thickness from 8 to 27 m(actual depositional cycles would have been 1 to 3 mthick), showing that arbitrary subdivision into layers ofequal thickness, as was done by McNamara and Ward-law (1991) in the analysis of Westerose, is not realistic.The lower Ricinus West reef interior appears to consist

of irregular, porous and permeable slices that rangefrom 15 to 30 m thick, and the available data suggestthat layers cannot be considered to be laterally continu-ous throughout the reservoir. The reef margin facies arethick and laterally discontinuous, with porous and per-meable zones mainly controlled by subvertical fracturesand brecciated intervals.

The distribution of porous and permeable zoneswithin the dolomitized Ricinus West buildup differs fromthat of the Westerose buildup updip in the central part ofthe Rimbey-Meadowbrook reef trend (McNamara and

Figure 19. Porosity and permeability histograms from tabular stromatoporoid boundstone facies inRicinus buildup margin well 7-13. Permeability increases slightly with increasing porosity (seeFigure 12 for vertical plots of these data). Porosity is slightly bimodal.

Page 309: Reservoir Quality Prediction in Sand and Carbonates

Wardlaw, 1991), where the highest porosity and perme-ability rocks occur in the lower portion of the reef interior.This appears to be due to differences in primary facies dis-tribution in the Westerose buildup because it is small (1.6× 5.5 km and 214 m thick) compared to the Ricinus West

(4 × 9.5 km and 259 m thick). However, the main differ-ences appear to be related in part to the differentapproaches used to map and group porosity and perme-ability data. In Ricinus West and the dolomitized part ofStrachan (D3A pool), distribution of pore types closelyfollows the depositional sequences (Figure 5).

Predicting Reservoir Quality

It is difficult to extrapolate reservoir quality beyondthe present database. However, comparing our datafrom the deep basin with information from intermedi-ate and shallower parts of the basin and with what wereonce more deeply buried strata in Rocky Mountain out-crops reveals several important aspects. Thus, somegeneralizations and predictions can be advanced bycomparing and integrating the above information withdata from limestone and dolomite reservoirs elsewherein the basin. Data from limestone buildups in the shal-lower parts of the Alberta basin suggest that the effectsof sea floor and early diagenesis are relatively minor,except for (1) some buildup margins that contain exten-sive submarine cements (e.g., Golden Spike) (Mountjoyand Walls, 1977; Walls et al., 1979) and (2) subaeriallyexposed portions of buildup interiors that are locallystrongly cemented (e.g., Golden Spike) (Walls and Bur-rowes, 1985, 1990). In the case of the older, isolatedSwan Hills buildups (Havard and Oldershaw, 1976;Wong and Oldershaw, 1981; Wendte and Muir, 1995),subaerial exposure also has had a comparatively minoreffect on reservoir quality; more than 90% of the poros-ity is of primary depositional origin.

The facies distributions in the buildups were pre-dominantly controlled by sea level changes and thusare widespread and generally predictable (Wendte,1992a, b), but with important variations caused by sed-iment supply and currents (McLean and Mountjoy,1993b). Subaerial unconformities formed during sealevel drops are also widespread and predictable(McLean and Mountjoy, 1994; Wendte and Muir,1995). It is reasonable to assume that these nonporouscemented zones also are present locally in the dolomi-tized buildups. This, together with detailed observa-tions from the partially dolomitized Miette buildupexposed in the Rocky Mountains (Mattes and Moun-tjoy, 1980) and reconnaissance observations from otherdolomitized Rocky Mountain buildups (McLean andMountjoy, 1993a, b, 1994), indicates that porousdolomites are regionally extensive and are largely con-trolled by the facies distribution and porosity of theoriginal limestones. Thus, porosity is primarily faciescontrolled and the Leduc limestone buildups such as Redwater (Klovan, 1964) and Golden Spike(McGillivray and Mountjoy, 1975; Walls, 1978) can beused as general guides to predict the types and distrib-ution of porosity and permeability, provided one takesinto account porosity reduction with increasing burial.

Predicting the distribution of replacement dolomitesand associated dissolution that took place during shal-low burial is also difficult. However, most of thebuildups in the Alberta basin are dolomitized, appar-ently because most were connected to conduit systemsin the underlying platforms (e.g., Rimbey-Meadow-brook reef trend) (Amthor et al., 1993; Mountjoy and

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 299

Figure 20. Crossplots of porosity and permeabilitydata (from Figure 19) from tabular stromatoporoidboundstone facies in Ricinus buildup margin well 7-13. The interval 4429 to 4440 m shows a clearer trendof permeability increasing with porosity than do theoverlying 4414 to 4418 or 4421 to 4424 m intervals.

Page 310: Reservoir Quality Prediction in Sand and Carbonates

300 Mountjoy and Marquez

Amthor, 1994) or were connected to fault and fractureconduit systems as in parts of the Peace River arch(Packard et al., 1990; Mountjoy and Halim-Dihardja,1991) and elsewhere (D. Green, 1996, personal communi-cation). Buildups away from these conduit systems suchas Golden Spike, Redwater, and Miette are not dolomi-tized, or only partially dolomitized, because apparentlythey were not linked to a regional conduit system at thetime of dolomitization. The four stages of porosity devel-opment relative to dolomitization (Figure 23) are repre-sentative of the diagenesis in different parts andstratigraphic levels of the basin (Walls and Burrowes,1985, 1990; Mountjoy, 1994; Mountjoy and Amthor, 1994).

Fracturing, except for late-stage microfracturing(Marquez and Mountjoy, 1996), has not been studiedsufficiently to determine its origin and overall distribu-tion. The somewhat greater abundance of fractures inthe buildup margins may be due to differential com-paction between the buildup and the adjacent basin andto the buildup margins being more strongly cemented.

With increasing burial, limestones gradually losetheir porosity due to pressure solution and cementa-tion, which in the Alberta Basin has reduced ordestroyed most of the primary porosity of limestonesburied deeper than 3500 m (Drivet, 1993; Amthor et al.,1994; Marquez, 1994; this study). Also, plugging of

pores and vugs in dolomite reservoirs by late-stageanhydrite, and to a minor extent carbonate cements, canlocally reduce porosity and permeability, as observed inthe Rimbey-Meadowbrook reef trend below depths of2300 m immediately updip from the Strachan reservoir(Drivet, 1993; Mountjoy et al., 1997). Bitumen pluggingcaused by deasphalting will take place in those reser-voirs in which crude oil has been cracked. Thus, exceptfor these important modifications, it is reasonable toinfer that dolomitized Leduc reservoirs elsewhere in thedeep basin will have porosity and permeability varia-tions similar to those observed in the Ricinus West andStrachan D3A reservoirs.

Although porosity and permeability are related todepositional facies and patterns of diagenetic overprint-ing, there is some variation within and between reefs, butthis depends on the scale. At the large scale (tens ofmeters), the porosity and permeability are remarkablyuniform (Figure 13); at the medium scale, the vertical andlateral arrangement of the dominant pore types variesconsiderably, causing variations in reservoir character. Atthe small (meter to millimeter) scale, porosity and perme-ability are controlled by facies, pore types, and differencesin diagenesis. The approach used for characterizing reser-voir properties and for classifying porosity, permeability,and pore systems is critical. Up to now, it has often been

Figure 21. Crossplots of porosity-permeability (log scales) of most-ly limestones in the Strachan D3A(A) and D3B (B) pools. Porosity scale is arithmetic and perme-ability scale is log. The two linesrepresent the 100-µm and 500-µmparticle size boundaries used byLucia (1983, 1995) to distinguishpermeability fields, in this caserock fabric/petrophysical class 1.Rock fabric/petrophysical class 2falls below the 100-µm line. TheStrachan limestones have lowporosities and permeabilities. Thehigher permeabilities from well 10-31 are due to permeability relat-ed to partial dolomitization.

Page 311: Reservoir Quality Prediction in Sand and Carbonates

one of arbitrary broad-scale lumping of horizontal slicesand blocks having similar porosities. A reservoir descrip-tion that closely follows the geology and depositionalfacies of the reservoir, and especially the diagenetic over-prints, as outlined here for the Ricinus West buildup, pro-vides a more realistic and accurate subdivision ofreservoir properties. This cannot be performed unless thedepositional facies, diagenesis, and pore types are thor-oughly documented by means of careful core observa-tions. In addition, grain and crystal size and sorting, andseparate/connected-vug type and porosity (Lucia, 1995),are important in terms of describing rock fabrics.

CONCLUSIONS

The association and distribution of pore types andpermeability within the Leduc Strachan and RicinusWest gas reservoirs in the deep Alberta basin indicate:

1. Depositional facies ultimately controlled the distri-bution of different pore types, whereas permeabil-ity is mainly controlled by diagenetic processes,especially dolomitization and dissolution, and var-ious phases of cementation in the lower part ofreservoirs below the paleo-oil/water contact and

bitumen in the upper part of reservoirs.2. Completely dolomitized Upper Devonian Leduc

buildups at depths >3000 m have higher porosi-ties and permeabilities than limestones becausedolostones are more resistant to pressure solu-tion during burial.

3. The relationships between the proportion ofdolomite and porosity are complex. In the Stra-chan buildup, a slight increase in porosity occurswith an increase in the amount of dolomite, whichbecomes more pronounced at >80% dolomite.This porosity increase appears to be related to theleaching of calcite and may be associated withdolomitization. In general, at burial depths >3000m, porosity and permeability increase withincreasing dolomitization, as in the partiallydolomitized Strachan buildup (D3A pool), withthe highest porosity and permeability occurringin completely dolomitized facies in the buildupinterior. At the Strachan reef margin, porous andpermeable dolostones are interbedded with non-porous and nonpermeable limestones, with stratathat had higher primary porosities being prefer-entially dolomitized. These trends are modifiedby later cementation, dissolution, and bitumenemplacement. At these depths, limestones make

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 301

Figure 22. Crossplots ofporosity-permeability (logscales) of the buildup inte-rior (A) and margin (B) inthe completely dolomitizedRicinus West reservoir.Lines and scales as inFigure 21. Data from well10-33 (A) includes intervalsplotted in Figure 11 andwell 7–13 in Figure 12. Theporosities and permeabili-ties are much higher thanthe values from limestonesin Figure 21. There is essen-tially no difference in thedistributions from thebuildup margin and interi-or. In general, the perme-abilities are higher thanexpected (compare withLucia, 1995) and appear tobe due to touching moldicvugs in a dolomite matrixwith good intercrystallineporosity (see text).

Page 312: Reservoir Quality Prediction in Sand and Carbonates

Stage 2: PARTIAL TO COMPLETE DOLOMITIZATION

A) Nucleation and early growth of dolomite rhombs in the lime matrix.

B) Further rhomb growth to form dolomite patches in lime matrix.

C) Advanced matrix replacement First dissolution of skeletal fragments and some lime matrix.

D) End of dolomite replacement. Partial to complete dissolution of calcite skeletal fragments.

REEF MARGIN REEF INTERIOR

Boundstones, rudstoneswith packstone matrices

Rudstones withpackstone andgrainstone matrices

Wackestones and mudstones

Stages 3 and 4: CEMENTATION, DISSOLUTION, MICROFRACTURING, AND BITUMEN PLUGGING

Lime mud, withminor calcite

Dolomite

Dolomiterhombohedrons

Dissolution vugsformed from calcite

Intercrystalline porosity

Stage 1: DEPOSITION

Submarinecements

Skeletalfragments

domal

tabular

Stromatoporoids

Increasing Dolom

itization

Figure 23. Inferred porosity evolution during replacement dolomitization of Leduc limestones and later diagene-sis illustrating changes and modification of pore types. Stage 1: deposition and early porosity reduction; Stage 2:replacement dolomitization; Stages 3 and 4: late cementation, dissolution, microfracturing, and bitumen plugging.

302 Mountjoy and Marquez

Page 313: Reservoir Quality Prediction in Sand and Carbonates

poor reservoirs because most of the primaryporosity has been filled with cement or destroyedby pressure solution.

4. Pore types closely follow depositional facies in theRicinus West buildup, despite complete dolomiti-zation. In general, porosity and permeability aresimilar throughout much of this buildup. At themeter scale, the upper buildup interior reservoir ischaracterized by 1- to 2-m-thick, permeable, later-ally continuous lagoonal zones separated bythicker intervals of nonporous, more finely crys-talline dolomite. The lower reef interior has dis-continuous permeable vuggy zones. The reefmargin contains intervals of breccia, fractures, andconnected vugs that provide vertical permeability.The best lateral permeability is related to laterallycontinuous skeletal wackestone to grainstonefacies that have been dolomitized to a rock withgood intercrystalline porosity in the upperbuildup interior. In the buildup margin, the bestvertical interconnections are provided by con-nected vugs and subvertical fractures in dolomiteswith intercrystalline porosity.

5. In dolomites, touching vugs of solution-enlargedmolds of Amphipora and stromatoporoids form anexcellent porous and permeable reservoir rock thatis related to primary reef margin and reef flatfacies. This moldic dolomite is an important reser-voir rock that, when surrounded by dolomiteswith intercrystalline porosity, has permeabilitieshigher than expected. These dolomites appear tobe largely responsible for the excellent productiv-ity of these reservoirs.

6. At the millimeter scale, porosity and permeabilityare controlled by diagenetic processes. Postdolomi-tization processes, such as pressure solution, cemen-tation, dissolution, and bitumen plugging, locallymodified porosity and permeability. Anhydritecementation, minor dolomite, calcite, and native sul-fur cementation reduced porosity and, more signifi-cantly, permeability, producing a high degree ofheterogeneity in the lower part of the reservoirs.

7. Bitumen reduces porosity and permeability bydecreasing pore and pore-throat sizes, mainly inthe upper part of the Ricinus West reservoir.Bitumen coating may cause wettability to changefrom water-wet to intermediate in water/gas sys-tems. Bitumen can only be determined by coreexamination, petrographic analysis, and coreporosity measurements. Errors in reservoir volu-metric calculations from well logs can arisebecause of the lack of significant density contrastbetween crude oil and bitumen, making bitumenindistinguishable from oil.

8. Predicting reservoir fabrics is difficult. However, inthis case, there is a general trend that Upper Devon-ian Leduc replacement dolomites form good toexcellent reservoir rocks in the deep Alberta Basin,except where plugged by late-stage calcite, anhy-drite, and bitumen. Early replacement dolomitiza-tion and associated calcite dissolution formed avuggy reservoir rock connected by a matrix ofintercrystalline porosity. The highest porosities andpermeabilities occur in those facies that had the

highest primary porosity and permeability; that is,the grainstones and framestones. The reservoircharacter was only slightly modified by later frac-turing and cement and bitumen fillings.

Reconstruction of the depositional and diagenetichistory from detailed studies of cores is essential formaking realistic reservoir models. Differences inapproach in grouping porosity and permeabilitydata in reservoir characterization can lead to majordifferences in modeling and assessment of a reser-voir. Critical to a realistic assessment is to describethe reservoir (depositional facies, sequences, anddiagenesis, including grain and crystal size, sorting,separate-vug type, and porosity) as geologicallyfaithfully as possible, rather than by means of arbi-trary slices or blocks. Such studies demonstrate thatcarbonate reservoirs are very heterogeneous, andthat discontinuous, relatively nonporous, and lowpermeability zones have the potential to impedefluid flow and greatly affect reservoir performanceand production. More realistic models are essentialfor understanding reservoir production and for pre-dictions of reservoir performance and ultimaterecovery by enhancement techniques. Only afterputting together such genuine reservoir models canthe more detailed petrophysical information beincorporated in them and assessed.

Although it is difficult to predict reservoir poros-ity and permeability trends beyond these two reser-voirs, it is clear that the secondary porosities in thesedeeply buried dolomites are mainly controlled by theprimary porosity distribution, which in turn is con-trolled by the depositional facies. Observations else-where in the deep basin and the adjacent RockyMountains suggest that these porous Leducdolomites are regionally extensive and should haveporosity and permeability variations similar to theStrachan and Ricinus West reservoirs.

ACKNOWLEDGMENTS

This research was supported by Natural Science andEngineering Research Council (NSERC) operating grantsand a strategic grant to E.W. Mountjoy and H.G. Machel,supplemented by funds from Chevron, Home Oil, MobilOil, Norcen, PanCanadian, Petro Canada, Shell Canada,and Imperial Oil. This paper has been modified andupdated from Marquez (1994). X. Marquez received ascholarship from Maraven S.A. and a grant-in-aid fromthe American Association of Petroleum Geologists. Weacknowledge the donation by Geographix of QLA2 loganalysis software, and Brett Norris for his assistance andinput with the log analysis. Individuals who shared theirwork and knowledge include: J. Amthor, G. Burrowes,G. Davies, W. Keith, B. Martindale, B. McNamara, H.Qing, J. Reimer, A. Rup, B. Scott, M. Teare, N. Wardlaw,B. Watt, and J. Wendte. Reviews of earlier versions andcomments by E. Drivet, J. Duggan, D. Green, H. Machel,J. Paquette, N. Wardlaw, and S. Whittaker improved themanuscript. We appreciate the helpful editorial com-ments and suggestions of Rick Major and John Bloch andthe helpful and thorough review by Julie Kupecz.

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 303

Page 314: Reservoir Quality Prediction in Sand and Carbonates

304 Mountjoy and Marquez

REFERENCES CITED

Amthor, J.E., E.W. Mountjoy, and H.G. Machel, 1993,Subsurface dolomites in Upper Devonian LeducFormation buildup, central part of Rimbey-Meadowbrook reef trend, Alberta, Canada: Bulletinof the Canadian Society of Petroleum Geologists, v. 41, p. 164–185.

Amthor, J.E., E.W. Mountjoy, and H.G. Machel, 1994,Regional-scale porosity and permeability variationsin Upper Devonian Leduc buildups: implicationsfor origin and distribution of porosity in carbonates:AAPG Bulletin, v. 78, p. 1541–1559.

Andrichuk, J.M., 1958a, Stratigraphy and facies analysisof Upper Devonian reefs in Leduc, Stettler and Red-water areas, Alberta: AAPG Bulletin, v. 42, p. 1–93.

Andrichuk, J.M., 1958b, Cooking Lake and DuvernayLate Devonian sedimentation in Edmonton area ofcentral Alberta, Canada: AAPG Bulletin, v. 42, p. 2189–2222.

Archie, G.E., 1952, Classification of carbonate reservoirrocks and petrophysical considerations: AAPG Bul-letin, v. 36, p. 278–298.

Barfoot, G.L., and R.J. Rodgers, 1984, Leduc, a new lifeat 38: Journal of Canadian Petroleum Technology,May–June, p. 41–46.

Barfoot, G.L., and S.C.M. Ko, 1987, Assessing, and com-pensating for, the impact of the Leduc D-3A gas capblowdown on the other Golden Trend pools: Journalof Canadian Petroleum Technology, July–August, p. 28–36.

Bateman, R.M., 1985, Open hole log analysis and forma-tion evaluation: Boston, IHRDC Publishers, 647 p.

Beales, F.W., and J.L. Hardy, 1980, Criteria for therecognition of diverse dolomite types with anemphasis on studies on host rocks for MississippiValley-type ore deposits, in D.H. Zenger, J.B. Dun-ham, and R.L. Ethington, eds., Concepts and mod-els of dolomitization: SEPM Special Publication 28, p. 197–213.

Burrowes, O.G., 1977, Sedimentation and diagenesis ofback-reef deposits, Miette and Golden Spikebuildups, Alberta: M.Sc. thesis, McGill University,Montreal, Quebec, 207 p.

Burrowes, O.G., and F.F. Krause, 1987, Overview ofthe Devonian System: subsurface Western CanadaBasin, in F.F. Krause and O.G. Burrowes, eds.,Devonian lithofacies and reservoir styles in Alberta:Calgary, 13th Canadian Society of Petroleum Geol-ogists Core Conference and Display, p. 1–20.

Carpenter S.J., and K.C. Lohmann, 1989, δ18O and δ13Cvariations in Late Devonian marine cements fromthe Golden Spike and Nevis reefs, Alberta, Canada:Journal of Sedimentary Petrology, v. 59, p. 792–814.

Choquette, P.W., A. Cox, and W.J. Meyers, 1992, Char-acteristics, distribution and origin of porosity inshelf dolostones: Burlington-Keokuk Formation:Journal of Sedimentary Petrology, v. 62, p. 167–189.

Choquette, P.W., and L.C. Pray, 1970, Geologic nomen-clature and classification of porosity in sedimentarycarbonates: AAPG Bulletin, v. 54, p. 207–250.

Chouinard, H., 1993, Lithofacies and diagenesis of the

Cooking Lake Platform carbonates, Alberta Basinsubsurface, Canada: M.Sc. thesis, McGill Univer-sity, Montreal, Quebec, 101 p.

Drivet, E., 1993, Diagenesis and reservoir characterizationof Upper Devonian Leduc dolostones, southern Rim-bey-Meadowbrook reef trend, central Alberta: M.Sc.thesis, McGill University, Montreal, Quebec, 115 p.

Drivet, E., and E.W. Mountjoy, 1993, Porosity variationsin Upper Devonian Leduc dolomites, central Rim-bey-Meadowbrook reef trend, Alberta (abs.): AAPGBulletin, Annual Convention Abstracts, p. 93.

Drivet, E., and E.W. Mountjoy, 1994, Timing ofdolomitization and secondary porosity in UpperDevonian Leduc dolostones, southern Rimbey-Meadowbrook reef trend, Alberta (abs.): CanadianSociety of Exploration Geophysicists and CanadianSociety of Petroleum Geologists Joint Annual Con-vention, Calgary, Alberta, p. 347–348.

Drivet, E., and E.W. Mountjoy, 1997, Burial dolomitiza-tion in the Leduc Formation (Upper Devonian),southern Rimbey-Meadowbrook reef trend, Alberta:Journal of Sedimentary Research, v. 67, p. 411–423.

Enos, P., and L.H. Sawatsky, 1981, Pore networks inHolocene carbonate sediments: Journal of Sedimen-tary Petrology, v. 51, p. 961–985.

Ghosh, S.K., and G.M. Friedman, 1989, Petrophysics of adolostone reservoir: San Andres Formation (Per-mian), West Texas: Carbonates and Evaporites, v. 4,p. 45–119.

Havard, C., and A. Oldershaw, 1976, Early diagenesis inback-reef sedimentary cycles, Snipe Lake reef com-plex, Alberta: Bulletin of Canadian Petroleum Geol-ogy, v. 24, p. 27–69.

Hriskevich, M.E., J.M. Faber, and J.R. Langton, 1980,Strachan and Ricinus West gas fields, Alberta,Canada, in M.T. Halbouty, ed., Giant oil and gasfields of the decade 1968–1978: AAPG Memoir 30,p. 315–328.

Hugo, K., 1990, Mechanisms of groundwater flow andoil migration associated with Leduc reefs: Bulletin ofthe Canadian Society of Petroleum Geologists, v. 38,p. 307–319.

Illing, L.V., 1959, Deposition and diagenesis of someUpper Paleozoic carbonate sediments in westernCanada: New York, Proceedings of the Fifth WorldPetroleum Congress, Section 1, p. 23–52.

Jardine, D., D.P. Andrews, J.W. Wishart, and J.W.Young, 1977, Distribution and continuity of carbon-ate reservoir: Journal of Petroleum Technology, v. 29,p. 873–885.

Jardine, D., and J.W. Wishart, 1982, Carbonate reser-voir description: Dallas, Texas, Society of PetroleumEngineers, Paper 10010, 13 p.

Klovan, J.E., 1964, Facies analysis of the Redwaterreef complex, Alberta, Canada: Bulletin of theCanadian Society of Petroleum Geologists, v. 12,p. 1–100.

Krouse, H.R., C.A. Viau, L.S. Eliuk, A. Ueda, and S.Halas, 1988, Chemical and isotopic evidence of ther-mochemical sulphate reduction by light hydrocarbongases in deep carbonate reservoirs: Nature, v. 333, p. 415–419.

Laflamme, A.K., 1990, Replacement dolomitization in

Page 315: Reservoir Quality Prediction in Sand and Carbonates

the Upper Devonian Leduc and Swan Hills forma-tions, Caroline area, Alberta, Canada: M.Sc. thesis,McGill University, 138 p.

Layer, D.B., 1949, Leduc oil field, Alberta, a Devoniancoral reef discovery: AAPG Bulletin, v. 33, p. 572–602.

Lishman, J.R., 1969, Core permeability anisotropy:Petroleum Society of the Canadian Institute of Min-ing and Metallurgy, 20th Annual Technical Meeting,Edmonton, Alberta, Paper 6920.

Lomando, A.J., 1992, The influence of solid reservoirbitumen on reservoir quality: AAPG Bulletin, v. 76,p. 1137–1152.

Lucia, F.J., 1983, Petrophysical parameters estimatedfrom visual descriptions of carbonate rocks: a fieldclassification of carbonate pore space: Journal ofPetroleum Technology, v. 35, p. 629–637.

Lucia, F.J., 1995, Rock-fabric/petrophysical classifica-tion of carbonate pore space for reservoir character-ization: AAPG Bulletin, v. 79, p. 1275–1300.

Lucia, F.J., and R.D. Conti, 1987, Rock fabric, perme-ability, and log relationships in an upward-shoal-ing, vuggy carbonate sequence: University of Texasat Austin, Bureau of Economic Geology GeologicalCircular 87–5, 22 p.

Machel, H.G., and E.W. Mountjoy, 1987, General con-straints on extensive pervasive dolomitization andtheir application to the Devonian carbonates ofwestern Canada: Bulletin of Canadian PetroleumGeology, v. 35, p. 143–158.

Maddox, D.F., 1984, Reservoir simulation study—Rici-nus West D-3A pool, Ricinus West field: CanterraEnergy Ltd. Internal Report, Calgary, Alberta, 90 p.

Marquez, X., 1994, Reservoir geology of UpperDevonian Leduc buildups, deep Alberta Basin:Ph.D. thesis, McGill University, Montreal, Quebec,285 p.

Marquez, X., and E.W. Mountjoy, 1996, Microfracturesdue to overpressures caused by thermal cracking inwell-sealed Upper Devonian reservoirs, deepAlberta Basin: AAPG Bulletin, v. 80, p. 570–588.

Mattes, B.W., and E.W. Mountjoy, 1980, Burialdolomitization of the Upper Devonian Miettebuildup, Alberta, in D.H. Zenger, J.B. Dunham, andR.L. Ethington, eds., Concepts and models ofdolomitization: SEPM Special Publication 28, p.259–297.

Mazzullo, S.L., 1992, Geochemical and neomorphicalteration of dolomite, a review: Carbonate andEvaporites, v. 7, p. 21–37.

McCaffery, F.G., 1977, Rock-fluid relationship studieson the Windfall D-3A reservoir and their applicationin evaluating gas cycling effectiveness: Journal ofCanadian Petroleum Technology, January–March,p. 55–63.

McGillivray, J.G., and E.W. Mountjoy, 1975, Facies andrelated reservoir characteristics, Golden Spike reefcomplex, Alberta: Bulletin of the Canadian Societyof Petroleum Geologists, v. 23, p. 753–809.

McLean, D.J., 1992, Upper Devonian buildup develop-ment in the Southern Canadian Rocky Mountains: asequence stratigraphic approach: Ph.D. thesis,McGill University, Montreal, Quebec, 290 p.

McLean, D.J., and E.W. Mountjoy, 1993a, Stratigraphy

and depositional history of the Burnt TimberEmbayment, Fairholme Complex, Alberta: Bulletinof Canadian Petroleum Geology, v. 41, p. 290–306.

McLean, D.J., and E.W. Mountjoy, 1993b, UpperDevonian buildup, margin and slope developmentin the southern Canadian Rocky Mountains: Geo-logical Society of America Bulletin, v. 105,p. 1263–1283.

McLean, D.J., and E.W. Mountjoy, 1994, Allocycliccontrol on Late Devonian buildup development,Southern Canadian Rocky Mountains: Journal ofSedimentary Research, v. B64, p. 326–340.

McNamara, L.B., and N.C. Wardlaw, 1991, Geologicaland statistical description of the Westerose reservoir,Alberta: Bulletin of Canadian Petroleum Geology, v. 39, p. 332–351.

McNamara, L.B., N.C. Wardlaw, and M. McKellar,1991, Assessment of porosity from outcrops ofvuggy carbonate and application to cores: Bulletinof Canadian Petroleum Geology, v. 39, p. 260–269.

Mossop, G.D., 1972, Origin of peripheral rim, Redwaterreef, Alberta: Bulletin of Canadian Petroleum Geol-ogy, v. 20, p. 238–280.

Mountjoy, E.W., 1994, Dolomitization and the characterof hydrocarbon reservoirs; Devonian of WesternCanada, in A. Parker and B. Sellwood, eds., Quanti-tative diagenesis, recent developments and applica-tions to reservoir geology: Dordrecht, KluwerAcademic Publishers, p. 33–94.

Mountjoy, E.W., and J.E. Amthor, 1994, Has burialdolomitization come of age? Some answers fromthe Western Canada Sedimentary Basin, in B.Purser, M. Tucker, and D. Zenger, eds., Dolomites,a volume in honour of Dolomieu: International Asso-ciation of Sedimentologists Special Publication 21,p. 203–229.

Mountjoy, E.W., and M.K. Halim-Dihardja, 1991, Multi-ple phase fracture and fault-controlled burial dolomi-tization, Upper Devonian Wabamun Group, Alberta:Journal of Sedimentary Petrology, v. 61, p. 590–612.

Mountjoy, E.W., H. Qing, E. Drivet, X. Marquez, S. Whit-taker, and A. Williams-Jones, 1997, Variable fluid andheat flow regimes in three Devonian dolomite con-duit systems, Western Canada Sedimentary Basin:isotopic and fluid inclusion evidence/constraints, inI.P. Montanez, J.M. Gregg, and K.L. Shelton, eds.,Basin wide fluid flow and diagenetic patterns: inte-grated petrologic, geochemical and hydrological con-siderations: SEPM Special Publication 57, p. 119–137.

Mountjoy, E.W., and R.A. Walls, 1977, Some examplesof early submarine cements from Devonianbuildups of Alberta: Miami, Rosentheil School ofMarine and Atmospheric Science, University ofMiami, Proceedings of the 3d International CoralReef Symposium, v. 2, p. 155–161.

Packard, J.J., G.J. Pellegrin, I.S. Al-Aasm, I. Samson,and J. Gagnon, 1990, Diagenesis and dolomitizationassociated with hydrothermal karst in Famennianupper Wabamun ramp sediments, northwesternAlberta, in G.R. Bloy and M.G. Hadley, eds., Thedevelopment of porosity in carbonate reservoirs:Canadian Society of Petroleum Geologists Continu-ing Education Short Course, Section 9, 25 p.

Podrusky, J.A., J.E. Barclay, A.P. Hamblin, L.P. Lee,

Predicting Reservoir Properties in Dolomites: Upper Devonian Leduc Buildups, Deep Alberta Basin 305

Page 316: Reservoir Quality Prediction in Sand and Carbonates

K.G. Osadetz, R.M. Procter, and G.C. Taylor, 1987,Conventional oil resources of Western Canada, PartI: Reservoir endowment: Geological Survey ofCanada, Paper 87–26, 42 p.

Reitzel, G.A., and G.O. Callow, 1977, Pool descriptionand performance analysis leads to understandingGolden Spikes miscible flood: Journal of PetroleumTechnology, v. 29, p. 867–872.

Reitzel, G.A., G. Davidson, G.O. Callow, and D.R. Bates,1976, Golden Spike D3-A pool oil depletion study,Calgary: Imperial Oil Ltd., Producing Department,Western Region, Report IPRC-5ME-76, 38 p.

Rogers, M.A., J.D. McAlary, and N.J. Bailey, 1974, Signif-icance of reservoir bitumens to thermal maturationstudies, Western Canada Basin: AAPG Bulletin, v. 58,p. 1806–1824.

Schmoker, J.W., and R.B. Halley, 1982, Carbonateporosity versus depth: a predictable relation forSouth Florida: AAPG Bulletin, v. 66, p. 2561–2570.

Seifert, S.R., 1990, Strachan Leduc gas pool, in M.L.Rose, ed., Oil and gas pools of Canada: CanadianSociety of Petroleum Geologists, v. 1, variously pag-inated.

Stoakes, F.A., 1992, Woodbend megasequence, in J.Wendte, F.A. Stoakes, and C.V.Campbell, eds.,Devonian–Early Mississippian carbonates of theWestern Canada Sedimentary Basin: a sequencestratigraphic framework: SEPM Short Course 28,Calgary, p. 183–206.

Van de Graff, W.J.E., and P.J. Ealey, 1989, Geologicalmodeling for simulation studies: AAPG Bulletin,v. 73, p. 1436–1444.

Vavra, C.L., M.H. Scheihing, and J.D. Klein, 1991,Reservoir geology of the Taylor sandstone in theOak Hill field, Rusk County, Texas: integration ofpetrology, sedimentology, and log analysis fordelineation of reservoir quality in a tight gas sand,in R. Sneider, W. Masssell, R. Mathis, D. Loren, andP. Wichmann, eds., The integration of geology, geo-physics, petrophysics and petroleum engineering inreservoir delineation, description and management:AAPG Proceedings of the 1st Archie Conference,Houston, Texas, p. 130–158.

Walls, R.A., 1978, Cementation history and porositydevelopment, Golden Spike Devonian reef complex,Alberta: Ph.D. thesis, McGill University, Montreal,Quebec, 307 p.

Walls, R.A., 1983, Golden Spike reef complex, Alberta,in P.A. Scholle, D.G. Bebout, and C.H. Moore, eds.,Carbonate depositional environments: AAPGMemoir 33, p. 445–453.

Walls, R.A., and O.G. Burrowes, 1985, The role ofcementation in the diagenetic history of Devonianreefs, western Canada, in N. Schneidermann andP.M. Harris, eds., Carbonate cement: SEPM SpecialPublication 36, p. 185–220.

Walls, R.A., and O.G. Burrowes, 1990, Diagenesis andreservoir development in Devonian limestone anddolostone reefs of western Canada: Canadian Soci-ety of Petroleum Geologists, Short Course Notes,Section 5, p. 5-1 to 5-18.

Walls, R.A., E.W., Mountjoy, and P. Fritz, 1979, Isotopic

composition and diagenetic history of carbonatecements in Devonian Golden Spike reef, Alberta:Geological Society of America Bulletin, v. 90, p. 963–982.

Wardlaw, N.C., 1980, The effects of pore structure ondisplacement efficiency in reservoir rocks and inglass micromodels: Society of Petroleum EngineersBulletin, no. 8843, p. 345–352.

Wardlaw, N.C., 1990, Characterization of carbonatereservoirs for enhanced oil recovery: Proceedings ofthe 1st Technical Symposium on Enhanced OilRecovery, Tripoli, Libya, Paper 90-01-05, p. 85–105.

Wardlaw, N.C., 1992, Effects of carbonate rock-poresystems on oil recovery, in Subsurface dissolutionporosity in carbonates: Recognition, causes andimplications: AAPG Short Course, Calgary, p. 1–28.

Waring, W.W., and D.B. Layer, 1950, Devoniandolomitized reef, D-3 reservoir, Leduc field,Alberta, Canada: AAPG Bulletin, v. 34, p. 295–312.

Weber, K.J., 1986, How heterogeneity affects oil recov-ery, in L.W. Lake and H.B. Carroll, eds., Reservoircharacterization: Orlando, Florida, Academic Press, p. 487–544.

Wendte, J.C., 1974, Sedimentation and diagenesis ofthe Cooking Lake platform and Lower Leduc reeffacies, Upper Devonian Redwater, Alberta: Ph.D.thesis, University of California, Santa Cruz, 222 p.

Wendte, J.C., 1992a, Cyclicity of Devonian strata in theWestern Canada Sedimentary Basin, in J. Wendte,F.A. Stoakes, and C.V. Campbell, eds., Devonian–Early Mississippian carbonates of the WesternCanada Sedimentary Basin: a sequence stratigraphicframework: SEPM Short Course 28, Calgary, p. 25–40.

Wendte, J.C., 1992b, Platform evolution and its controlon reef inception and localization, in J. Wendte,F.A. Stoakes, and C.V. Campbell, eds., Devonian–Early Mississippian carbonates of the WesternCanada Sedimentary Basin: a sequence stratigraphicframework: SEPM Short Course 28, Calgary, p. 41–88.

Wendte, J.C., 1994, Cooking Lake Platform evolutionand its control on Late Devonian Leduc reef incep-tion and localization, Redwater, Alberta: Bulletin ofCanadian Petroleum Geology, v. 42, p. 499–528.

Wendte, J.C., and I. Muir, 1995, Recognition and signifi-cance of an intraformational unconformity in LateDevonian Swan Hills reef complexes, Alberta, in D.A.Budd, A.H. Saller, and P.M. Harris, eds., Unconformi-ties and porosity in carbonate strata: AAPG Memoir63, p. 259–278.

Wendte, J.C., and F.A. Stoakes, 1982, Evolution and cor-responding porosity distribution of the Judy Creekreef complex, central Alberta, in W.G. Cutter, ed.,Canada’s giant hydrocarbon reservoirs: CanadianSociety of Petroleum Geologists, Core ConferenceManual, CSPG-AAPG Convention, Calgary, Alberta,p. 63–81.

Wong, P.K., and A. Oldershaw, 1981, Burial cementa-tion in the Devonian Kaybob reef complex, Alberta,Canada: Journal of Sedimentary Petrology, v. 51, p. 507–520

306 Mountjoy and Marquez