Anda di halaman 1dari 22

1

Grammer, G. M., P. M. Harris, and G. P. Eberli, 2004, Integration of outcrop and


modern analogs in reservoir modeling: Overview with examples from the
Bahamas, in Integration of outcrop and modern analogs in reservoir modeling:
AAPG Memoir 80, p. 1 22.

Integration of Outcrop and Modern


Analogs in Reservoir Modeling: Overview
with Examples from the Bahamas
G. Michael Grammer
Western Michigan University, Department of Geosciences, Kalamazoo, Michigan, U.S.A.

Paul M. Mitch Harris


ChevronTexaco, Energy Technology Company, San Ramon, California, U.S.A.

Gregor P. Eberli
University of Miami, Comparative Sedimentology Laboratory, Miami, Florida, U.S.A.

INTRODUCTION

was done previously with static depositional or facies


models. These dynamic conceptual models offer better
predictability of the distribution of potential reservoir
facies and their reservoir quality, especially when combined with recent advances in our understanding of
the detailed internal architecture and diagenesis of depositional systems.
In many reservoirs, key components that must be
well understood for reservoir characterization include
the facies types and probable diagenetic alteration, as
well as the vertical stacking patterns associated with
high-frequency cycles. It is at this scale (i.e., meter scale)
that fundamental controls over fluid-flow behavior in
the reservoir are commonly exerted. Thorough understanding of the nature of these cycles in the subsurface,
however, can be challenging, especially in cases where
core control is less than optimal. It is here that outcrop
analogs have been shown to provide valuable insight
into the understanding of the boundaries and internal
facies distribution of these high-frequency cycles.
Rarely, however, do we get a full three-dimensional
(3-D) picture of potential reservoir distribution from
outcrop alone, especially from the standpoint of aerial
dimensions. Outcrop studies typically provide solid data
on the vertical dimensions of a system, but the lateral
aspects are generally limited because the exposures are
either strike oriented, dip oriented, or at some orientation oblique to these directions. As a means to enhance

Development of a geologically constrained reservoir


model and subsequent upscaling of the model for reservoir simulation depends on critical input parameters
defining both the geometrical attributes and distribution of the targeted reservoir facies. To accurately characterize the potential reservoir, one must address the
geologically defined variability in the system. Gross differences in sedimentary facies, as well as more local variations in aspects such as grain size/type, grain sorting,
sedimentary structures, and diagenetic overprint, may
all influence the internal makeup and geometry of sedimentary deposits and, thus, the heterogeneity of potential reservoirs. Integration of geologically based elements
is, therefore, a fundamental step in the characterization
of the probable lateral and vertical distribution and
variability of reservoir facies in the subsurface. Such a
geologically based model not only increases our understanding of reservoir heterogeneity but also provides
the foundation for which the rest of the reservoir models and, ultimately, simulation models can be built.
In the last several years, we have seen the development of high-resolution sequence and cycle stratigraphy,
and with it, the advent of a refined mode of interpretation
for depositional systems. Using a sequence-stratigraphic
approach, sedimentary systems are analyzed dynamically through time, rather than as a single time slice, as

Grammer et al.

our understanding of the two-dimensional (2-D) aerial


distribution of possible reservoir facies, it can be invaluable to incorporate spatial data from the study of modern analogs (e.g., Grammer et al., 2001). Through the
evaluation of modern depositional systems, we can realize a first-order approximation of the 2-D aerial distribution of principal facies belts in a particular depositional setting during at least one time slice (i.e., the
present Holocene). There are undoubtedly limitations
to using the modern as an analog because certain
boundary conditions, such as climate and tectonic variability, relative position of sea level, and rates of sea
level rise or fall, might be different in ancient systems
being studied. Nevertheless, in the absence of highquality 3-D seismic data, modern depositional systems
are one of the only data sets from which reservoir modelers can extract the 2-D aerial distribution of potential
reservoir facies that is needed to accurately drive geologically constrained 3-D reservoir models.

OUTCROP ANALOGS
AND LIMITATIONS
Outcrop models have typically been relied on to
generate a proxy for subsurface reservoir distribution
(e.g., Borer and Harris, 1991; Tyler and Finley, 1991;
Grant et al., 1994; Kerans et al., 1994; Grammer et al.,
1996; Tinker, 1996; Kerans and Tinker, 1997; Sarg et al.,
1999; White and Barton, 1999; McLaurin and Steele,
2000; Willis and White, 2000; Willis and Gabel, 2001;
Camacho et al., 2002). Although detailed outcrop models are invaluable for establishing high-resolution subsurface models, outcrop data are commonly somewhat
constrained by the limits of exposure.
Reservoirs in the San Andres/Grayburg Formations
of the United States Permian Basin illustrate the degree
of heterogeneity that can typically occur and the value
of outcrop analogs to address some key reservoir issues.
Permeability-thickness, pressure-contour, and recovery maps for San Andres/Grayburg reservoirs suggest a
strong aerial and vertical compartmentalization of flow
units in these reservoirs (Dulaney and Hadik, 1990;
Harris and Walker, 1990; Lucia et al., 1990; Major et al.,
1990; Purves, 1990; Ruppel, 1990). Low field recoveries
are in stark contrast to laboratory core-flood experiments
that commonly indicate waterflood displacement efficiencies of as much as 50% original oil in place (OOIP),
depending on the wettability and pore structures of
the various rock types investigated. This disparity in
laboratory and field-recovery efficiency, despite a high
density of wells, emphasizes the need for better characterization of flow properties and remaining fluid
distribution in the reservoir. Flow modeling on a wellcharacterized system of properties distributed in a geo-

logic framework will lend some insight into the observed differences in laboratory and field recoveries.
Whenever these differences cannot be explained, both
the distribution of flow properties and the flow simulation should be reexamined.
As an example, outcrop study of the Permian San
Andres Formation along the Algerita Escarpment of the
Guadalupe Mountains has shown that the San Andres
is highly cyclical and that the distribution of these cycles
may influence both vertical and lateral reservoir properties. The cycles in the San Andres are typically 3 12
m thick, with facies formed during relative rise and/or
stabilization of sea level, during which carbonate sand
shoals developed. The cycles consist of thin mudstone/
wackestone bases, overlain by burrowed wackestones
and packstones, and capped by thick massive to planar
or cross-bedded packstones and grainstones (Kerans
et al., 1994), indicating the potential for vertical compartmentalization. The outcrops also display lateral facies relationships in the cycles on the scale of hundreds
of meters, which are representative of those commonly
observed in analogous hydrocarbon reservoirs of the
Permian Basin.
Outcrop analog models of ramp-carbonate reservoir
heterogeneity (e.g., Grant et al., 1994; Kerans et al.,
1994) may provide some insight into subsurface recovery histories. At a gross scale, stratigraphic relationships can be used to predict the occurrence of welldeveloped cycles that may contribute the most to production during the earliest phases of a waterflood.
Conversely, poorly developed cycles in thin-bedded
and massive fusulinid zones, which retain most of the
bypassed oil and cause flow to be compartmentalized,
also can be predicted by a better understanding of stratigraphic relationships. Potential variations between the
outcrop analogs and reservoirs themselves, however,
occur for the following reasons: (1) aerial permeability
variations will cause significant reductions in the estimated aerial sweep; (2) permeability may be correlated
better laterally in discrete stratigraphic horizons in the
subsurface than observed in outcrop; and (3) significant
differences exist in average permeability in cycles, resulting in increased channeling of injected fluids and
decreased overall vertical sweep efficiency.
An important point to consider in either an outcrop
or reservoir study is the vertical scale of the various
stratigraphic units, i.e., the vertical scale of layers to be
used in a reservoir model. Cycles, groups of cycles, and
less distinctly cyclic zones are geologic layers that have
potential significance from a reservoir perspective. They
can be recognized in core and log studies and should
form the basis of a geologic framework in a subsurface
model. Once these layers are identified and correlated,
petrophysical analyses of lithofacies in individual cycles,
or less distinctly cyclic units, yield the basic data needed
to model and predict the timing and efficiency of oil

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

recovery for these zones. This assumes, of course, that


the lateral correlation of permeability is small relative
to the interwell distance.
Outcrop analogs provide 2-D and occasional 3-D
views of facies with better details of diagenetic overprint
than is found in the subsurface or in modern analogs.
Porosity and permeability, when measured on outcrop
in a cycle/sequence and facies framework, are used to
illustrate possible fluid-flow scenarios in reservoir and
simulation models. As such, outcrop analogs provide a
more complete view of both facies and reservoir quality
dimensions than can be gleaned from subsurface data
alone. Although stratigraphic and facies relationships
recognized on outcrop can be useful as a template for
correlating subsurface data, questions always remain regarding how well the details of facies and diagenesis
from an outcrop actually compare with a particular reservoir or reservoir layer.

MODERN ANALOGS
AND LIMITATIONS
Modern environments are valuable as analogs for
conceptualizing the spatial distribution of reservoir facies in a single time slice and for obtaining a first-order
quantitative approximation of geometrical attributes
for potential reservoir facies. Combining remote sensing data, such as satellite images and aerial photographs,
with surface sediment maps provides facies trends and
dimensionality data that can be used to show patterns
and assist the modeling of a reservoir relative to simulated well spacing. These 2-D data are of even more value
when combined with results of coring studies, either
from modern environments where the spatial distribution of facies in a single depositional cycle is documented, or from detailed outcrop work that provides a more
comprehensive view of the third (i.e., vertical) dimension to the reservoir system. As with outcrop examples,
modern analogs have shortcomings when data derived
from them are to be applied to the subsurface. For example, the question of how well the facies patterns from
a Holocene example actually compare with a particular
reservoir or reservoir layer is always a cause for concern
and, therefore, a shortcoming. In addition, Holocene
studies are commonly insufficient analogs for portraying reservoir quality variations because of their limited
stratigraphic thickness and lack of diagenetic complexity. Many variables, such as variations in climate and
tectonic setting, relative variation in amplitude and
frequency of sea level changes, and, for carbonates in
particular, age-dependent faunal variability that may
result in changes in depositional architecture coupled
with complex diagenetic alteration are all limiting factors when incorporating modern analogs into a reservoir
model. Nevertheless, modern analogs have served as

primary models for many depositional systems over


the past several decades (e.g., Illing, 1954; Newell and
Rigby, 1957; Off, 1963; Purdy, 1963; Ball, 1967; Shinn
et al., 1969; Weil et al., 1973; Hardie, 1977; Huthnance,
1982; Amos and King, 1984; Eberli and Ginsburg, 1987;
McBride and Moslow, 1991; Davis and Balson, 1992;
Grammer and Ginsburg, 1992; Hulscher et al., 1993;
Major et al., 1996; Dyer and Huntley, 1999; Grammer
et al., 2001). New studies continue to further define the
spatial distribution of facies in the context of sea level
variability, which, when combined with study of early
diagenetic change in carbonates (e.g., Harris, 1978; Halley and Harris, 1979; Beach, 1993; Grammer et al.,
1993b; Melim et al., 1995; Anselmetti and Eberli, 1997;
Grammer et al., 1999), can provide invaluable information for 3-D reservoir characterization and modeling.

MODERN ANALOGS: EXAMPLES


FROM THE BAHAMAS
Studies of Great Bahama Bank (GBB), the largest
isolated carbonate platform in the Bahamas, continue to
aid in the refinement of stratigraphic, depositional, and
diagenetic models for carbonate systems. Models from
the Bahamas are of particular importance in understanding the architecture and reservoir quality of reservoirs in
isolated carbonate platforms.

Historical Perspective
Understanding the various carbonate facies types,
the controls on their formation, and their distribution
on carbonate platforms is, in large part, the result of
extensive study of modern analogs. Shortly after World
War II, the study of modern sedimentary environments
in carbonate (and siliciclastic) systems grew rapidly as
petroleum companies realized the value of applying
lessons from modern environments to similar ancient
reservoirs found in the subsurface. For carbonates, much
of this early study focused on the Bahamas, a broad expanse of modern carbonate banks southeast of the continental United States (Figure 1).
Studies in the Bahamas over the last several decades
have led to development of depositional facies models
for several settings, including platform interior, carbonate sand, reef, tidal flat, and marginal slope deposits
(see, for example, Illing, 1954; Newell and Rigby, 1957;
Purdy, 1963; Ball, 1967; Shinn et al., 1969; Ginsburg
and James, 1974; Hardie, 1977; Hine and Neumann,
1977; Beach and Ginsburg, 1980; Hine et al., 1981; Halley et al., 1983; Harris, 1983; Eberli and Ginsburg, 1987;
Grammer et al., 1993a; Major et al., 1996). In addition, significant insight has been gained about various
diagenetic processes and mechanisms in carbonates,

Grammer et al.

FIGURE 1. Location map for Great Bahama Bank (GBB), the Joulters Cays ooid shoal complex, and the Tongue of the
Ocean. Great Bahama Bank is separated from Florida by the Straits of Florida, through which the Gulf Stream flows.
Note also locations for the Western Line seismic data, the boreholes Clino and Unda, and the Ocean Drilling Program
drill sites.
including how they are recognized, as well as how they
may affect sediments from the standpoint of reservoir
potential (e.g., Harris, 1978; Halley and Harris, 1979;
Dawans and Swart, 1988; Anselmetti and Eberli, 1993,
1997; Beach, 1993; Grammer et al., 1993b, 1999; Melim
et al., 1995, 2002). Each of these studies has led to
further understanding of carbonate facies distribution
and early diagenetic modification that can enhance
both exploration and production strategies in subsurface reservoirs.
Recent stratigraphic and depositional studies of
GBB provide a fundamental understanding of the lateral growth potential and pulsed progradation of carbonate platforms. From seismic data, it is clear that
older isolated platforms coalesced to form GBB through
progradation along their leeward margins as a result
of highstand shedding of bank-top derived sediment
(Eberli and Ginsburg, 1987). The growth and diagenesis of platform strata are intimately linked to sea level
rise and fall, and on the platform top, the role of antecedent topography in initiating development of both
marginal reefs and sand bodies is strongly coupled to
the windward or leeward orientation of the platform
margin (Ball, 1967; Hine and Neumann, 1977; Harris,
1979; Hine et al., 1981; Halley et al., 1983). Likewise,
in the proximal slope environment, the sedimentary
makeup of facies (i.e., grain vs. mud-dominated) is de-

pendent on the windward/leeward orientation of the


margin (Grammer, 1991). Details of the genesis of platform-top shallowing-upward cycles on GBB, coupled
with realization that unfilled accommodation space is
common in both the Holocene and Pleistocene records,
also add to our understanding of equivalent strata in
ancient platforms. The nature of cycle variability in
GBB suggests limitations inherent in cyclostratigraphic
correlation and helps to explain aspects of reservoir heterogeneity that may be applicable to some ancient carbonate systems (Grammer et al., 2001).
In this chapter, several examples from the study of
modern environments in the Bahamas will be briefly
examined to illustrate the intrinsic value of incorporating data from modern depositional environments.
First, continuing studies of a platform-margin ooid
shoal environment that was first studied in detail in
the 1970s has provided details on the lateral heterogeneity of potential reservoirs at both exploration and
production scales. Lessons learned from these studies
have been incorporated into several studies of ancient
carbonate sand bodies. Next, studies initiated in the
early 1990s on the timing, evolution, and potential
reservoir distribution of marginal slope deposits in the
Bahamas again illustrate the value of incorporating
geometrical data from modern analogs. Following these
two examples, the chapter continues with a discussion

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

of some recent advances in our understanding of the


chronostratigraphic significance of seismic reflector
horizons and, finally, to a review of some of the major
controls on the petrophysical properties of carbonate
sediments and rocks. A brief synopsis of what the
authors believe are some of the key lessons learned
from the study of the Bahamas over the past couple of
decades is included in Table 1 and should leave the

reader with an appreciation for the value of comparative sedimentology in reservoir analysis.

Platform-margin Ooid Shoal


The Joulters Cays ooid shoal complex, located north
of Andros Island on GBB (Figure 1), exhibits spatial variability in both depositional and diagenetic facies that

Table 1. Key lessons learned from the Bahamas.


Platform evolution and stratigraphy
 Isolated carbonate platforms may coalesce through lateral progradation by highstand shedding of bank-top derived
sediment, especially along leeward margins.
 Highstand shedding of platform-derived sediment results in a highstand wedge of sediment on the slope that mimics a
lowstand wedge in both geometrical attributes and positioning.
 Marginal slope sedimentary processes and lithofacies are variable between windward and leeward margins.
 Failure of platform margins and slopes and the formation of megabreccias occurs during both lowstands and highstands.
 Biostratigraphy and magnetostratigraphy confirm that several of the regionally correlatable sequence boundaries are
chronostratigraphic horizons.
 Seismic reflector horizons in pure carbonate systems are typically the combined result of lithologic and diagenetic change.
 Unfilled accommodation space has significant implications to the application of cyclostratigraphy and the limitations
inherent in regional correlations.
Lithofacies and depositional bodies
 Lithofacies types and their distribution are relatively consistent across carbonate platforms and are a function of
paleogeography and paleoceanography (e.g., windward vs. leeward orientation, tidal range and influence, wave regimes,
paleolatitude, etc.).
 Ooid shoal and reefal belts tend to be best developed along windward margins, where they typically develop on preexisting
topographic highs.
 Whereas ooid shoal belts tend to initiate on antecedent highs, the resulting facies distribution and geometry is a result of
varying syndepositional processes.
 Sedimentary wedges along leeward slopes range from grain dominated to mud dominated, depending on the configuration
of the platform top.
 The genesis and evolution of shallowing-upward cycles have been defined in several environments (sand shoals, tidal flats,
interior platform, and reefs).
Diagenesis
 Diagenetic processes have been characterized in the freshwater vadose, meteoric phreatic, marine phreatic, and burial
diagenetic environments.
 Marine cementation occurs in shallow to very deep settings (platform top to basin floor).
 Syndepositional marine cementation occurs rapidly in the intertidal, shallow subtidal, and deeper subtidal (100 m) realms
and may influence the sedimentary deposits in a geologically instantaneous time frame.
 Depositional cycles influence the diagenetic overprint and subsequent reservoir quality in carbonates.
 Dissolution of aragonite grains in the marine phreatic zone, as well as the freshwater phreatic zone, can create moldic porosity.
 Dissolution and cementation processes continuously alter the physical properties of carbonates.
Inferences for reservoir distribution
 The distribution and nature of stratigraphic traps have been established in several environments (sand shoals, tidal flats,
interior platform, reefs, and slopes).
 Tidal flat, ooid shoal, reefal, and proximal slope reservoirs are laterally extensive along strike, but may be foreshortened in a
dip direction.
 Coarse-grained talus wedges with good initial reservoir potential form along proximal slopes during lowstand and
transgressive phases.
 Tidal flat, ooid shoal, and reefal reservoirs may be segregated horizontally and vertically at the production and enhancedproduction scales.

Grammer et al.

could ultimately lead to the formation of reservoir heterogeneity and the development of stratigraphic traps.
These carbonate sand bodies have been extensively
studied through surface sampling and shallow coring
(Harris, 1979, 1983, 1984) and have provided much of
our understanding of how ooid shoals develop and evolve
with changing sea level.
The Joulters example illustrates some of the difficulties associated with interpretation and correlation of
grainstones in subsurface studies of platform carbonate
reservoirs. The modern shoal complex, which extends
more than 400 km2, varies greatly in both thickness
and primary depositional fabric (Harris, 1979, 1983).
Antecedent topography and subsequent carbonate sand
generation varied greatly in the Joulters Cays area, both
during flooding of the platform and throughout development of the shoal complex during the latest rise of
sea level. Shoal growth, largely in response to the relative rise of sea level, records rapid expansion of ooid
sand belts, island formation and associated meteoric
diagenesis, development of marine hardgrounds, and
local shoal stabilization and reworking by burrowing
organisms.

A characteristic vertical succession revealed by coring the Joulters shoal (Figure 2) consists of scattered
lithoclast sands and pellet muds at the base, peloid sands
in the middle, and ooid sands at the top, showing an
upward increase in grain size, percentage of ooids, and
grain-supported fabric (Harris, 1979, 1984). This facies
sequence thins to the south over a shallowing Pleistocene surface and to the north and west as overall
sediment thickness decreases. Within the shoal complex, the thicknesses of the dominant facies are complementary; ooid sands thin in a bankward direction
as peloid sands thicken to form the thickest part of an
interplatform sheet. These facies changes resulted from
changing depositional patterns in response to rising
sea level (Harris, 1979). The shoal grew in three stages
(Figure 3): (1) an early bank-flooding stage, in which
muddy sands of peloids and pellets accumulated in
protected lows on the Pleistocene floor; (2) a shoalforming stage, during which ooid production began on
bedrock highs where bottom agitation was focused;
and (3) a stage of shoal development in which the production and dispersal of ooid sands established the
present size and physiography of the shoal.

FIGURE 2. Fence diagram modified from Harris (1979) showing the subsurface facies relations at the Joulters shoal complex
documented with extensive coring. Thin-section photomicrographs illustrate modern sediment equivalents of (A) ooid
grainstone; (B) ooid packstone; (C) peloid packstone; and (D) peloid wackestone. Photos are in plane polarized light.

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

FIGURE 3. Time-slice maps modified from Harris (1979) showing the complicated filling of accommodation space, both
regionally and locally, in the Joulters Cays area. With the rise of sea level during the Holocene, the Pleistocene limestone surface (shown in red at T-1) is quickly flooded. T-2: Burrowed muddy sands of peloids and skeletal fragments form
across the platform (orange), ooid sands (yellow) are localized on Pleistocene highs, and skeletal sands with reefs form
nearer to the platform margin (green). T-3: The ooid sand factory increases in size and dip width with continued sea level
rise; ooids are shown in the accompanying photomicrograph. T-4: Ooid sands have caught up to sea level over a broad
area, forming a vast, shallow, sand-flat setting (blue) and restricting the active ooid formation to a narrow belt along the
windward-facing sides of the shoal complex (yellow). The finer-grained nature of the sand-flat sediments is shown in the
accompanying photomicrograph.

Heterogeneity of potential reservoir quality in the


Joulters Cays shoal complex is inferred on the basis of
the distribution of depositional facies (Major et al.,
1996). Clean ooid sand along the active margin of the
shoal is deposited as subtidal-bar, channel-fill, beach,
and island facies. In cross section, the sand occurs as an
irregularly shaped area 2 km wide and 2 3 m thick.
High initial porosity values were measured in similar
clean carbonate sands by Halley and Harris (1979) and
Enos and Sawatsky (1981) and were confirmed in these
sands by thin-section analysis. Immediately bankward
of the clean ooid sand are widespread, somewhat irregularly shaped layers containing mixtures of carbonate
sand and mud that would likely result in vastly different reservoir properties in the subsurface. As discussed
by Major et al. (1996), there is an upper layer of muddy
ooid sand, some 20 km wide and from 4 to less than 1 m

in thickness that thins bankward and overlies a more


widespread lower layer of muddy, fine-grained peloid
sand. The upper ooid-rich layer would most likely have
better reservoir quality than the lower layer, at least initially, because of the larger grain size and lower overall
mud content.
Patterns of heterogeneity on a scale of hundreds of
meters (i.e., interwell scale) in the active part of the
Joulters Cays shoal also can be inferred from the facies
distribution (Major et al., 1996). A well-sorted ooid sand
facies occurs in the center of the active shoal, with a
more poorly sorted ooid sand facies both bankward
and seaward. Heterogeneity is inferred because of the
mud content, burrowing, and grain-type variations.
By analogy, similar subtle textural variations can be expected to produce local heterogeneity in ooid grainstone
reservoirs.

Grammer et al.

Earlier studies of both


modern and ancient carbonate slopes suggested many
downslope, gravity-induced
mechanisms for the deposition of sand-sized and coarsergrained sediments, most of
which suggested a relatively
large-scale (hundreds of meters to kilometers) distribution of coeval slope facies.
Several studies of modern
slopes have indicated that
relatively large-scale turbidity currents and debris flows
appear to be the dominant
mechanism for the downslope transport of coarse detritus (e.g., Cook and Mullins, 1983; Enos and Moore,
1983). On ancient carbonate
slopes, all types of sediment
gravity flows have been proFIGURE 4. Block diagram modified from Grammer and Ginsburg (1992) illustrating
posed, but again, the predomthe present-day surface characteristics and inferred three-dimensional structure of the inant depositional mechTongue of the Ocean foreslopes. The margin is characterized by a near-vertical escarp- anisms are interpreted to be
ment that formed as a result of erosional processes during multiple sea level fluctuations
debris flows or turbidity curin the Pleistocene Holocene. Erosion was augmented by some constructive growth of
rents (Cook, 1983).
organisms, resulting in slope angles for the escarpment that are typically 70 908. The
The more recent Bahasteeply dipping (35 458) slope deposits shown in light purple were deposited during
the lowstand and the early rise of sea level following the last major lowstand. These mas studies used manned
submersibles and state-ofunits reflect deposition from fringing reefs growing along the escarpment and have
the-art sampling and analythigh initial porosities. The downslope, highstand wedge of sediment is comprised of
sediment derived from the platform top. This wedge of sediment, which is as much as ical techniques to evaluate
90 m thick, was deposited during the last 7000 yr or so when the platform top was fully depositional and diagenetic
flooded. See Grammer and Ginsburg (1992) and Grammer et al. (1993a) for further
processes and their resultdiscussion.
ing products along marginal slopes in the Tongue of
the
Ocean
(Figure
1),
a
deep
intraplatform basin in the
Marginal Slope Environment
Bahamian archipelago (Ginsburg et al., 1991; GramAs with the discussion of carbonate sand bodies,
mer, 1991; Grammer and Ginsburg, 1992; Grammer et
studies of modern carbonate slopes can provide valal., 1993a). These studies suggest that at least some
uable insight into the sedimentary processes, deposisteep marginal slopes formed by alternative mechational architecture, and sequence evolution of equivnisms to those previously suggested. For example, dealent subsurface slope deposits. In these environments,
tailed analysis of the 35 458 marginal slope deposits
reservoir potential results from the interplay between
suggests that deposition took place through a combipaleogeographic orientation and geometry of the platnation of episodic rockfall and grain-flow processes on
form, mass-flow depositional processes, and early diaa more local scale, instead of through large-scale massgenetic modification. In addition, because sequenceflow events. The resulting deposits (Figures 4, 5) are charstratigraphic interpretations of carbonate platform maracterized by a series of dip-oriented lenses of relatively
gins are based to a large degree on concepts of variable
coarse debris that were deposited during transgression
timing and nature of deposition relative to fluctuaand early highstand (Grammer and Ginsburg, 1992).
tions in sea level, Quaternary platform margins, such
They are commonly characterized by a series of highly
as those found in the Bahamas, provide an opportuporous and permeable lenticular beds of coarse-grained
nity to calibrate the sedimentary record because of the
and poorly sorted sediments that are discontinuous in
well-constrained nature of sea level history during this
both strike and dip directions. These proximal slope
period.
deposits, which may form along either windward or

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

FIGURE 5. Representative photographs from steeply dipping marginal slope environment: (A) lenticular bedding that
makes up the 35 458 slope deposits. Beds are a few decimeters to half a meter in thickness, extend downslope for tens of
meters, and are characterized by coarse-grained skeletal fragments and talus with high initial porosities; (B) thin-section
photomicrograph (plane polarized light) of skeletal grainstone from the interior of the steeply dipping slope deposits
with much of the original porosity occluded by bladed Mg-calcite cement (M); (C) thin-section photomicrograph
(crossed nichols) showing compound growth of pore-filling botryoidal aragonite (BA). Early and pervasive marine
cementation by aragonite and Mg-calcite tends to occlude much of the high initial porosities in these deposits; (D) SEM
photomicrograph of fibrous aragonite precipitated in 8 months at a depth of 100 ft (30 m) along the platform margin in
the Bahamas (see Grammer et al., 1999 for details); (E) SEM photomicrograph of fibrous aragonite precipitated in 20
months at a depth of 100 ft (30 m) along platform margin showing a well-developed isopachous rim of marine cement.
Results of analytical and in-situ experiments (Grammer et al., 1993b, 1999) indicate that marine cements may partially
occlude carbonate sediments at depths of 30 75 m in just a few months.

10

Grammer et al.

leeward margins, have high initial reservoir potential,


but may be adversely affected by extensive syndepositional marine cementation (Grammer et al., 1993b,
1999).
Individual lenticular beds that make up the steep
marginal slopes in the Bahamas are only decimeters
thick and extend downslope for tens of meters and
would be expected to exhibit preferential flow characteristics developed in a dip dimension (Grammer et al.,
1993a). Multiple episodes of deposition during sea level
rise led to the amalgamation of these lenses, with the
result that the steeply dipping marginal slope environment consists of a broad apron of coarse-grained deposits that extends for tens of kilometers in a strike orientation but would be expected to exhibit rapid lateral
and vertical heterogeneity, in both strike and dip directions, at the meter and tens-of-meters scale.
In the distal slope environment, reservoir potential
is strongly dependent on the timing of sedimentation
(i.e., highstand vs. lowstand), as well as the windward
or leeward orientation of the margin. Extensive off-bank
shedding of fine-grained sediment, especially along leeward margins, results in rapid lateral progradation of
the platform (Wilber et al., 1990; Grammer and Ginsburg, 1992; Eberli et al., 1997). At present, downslope
transport of this fine-grained sediment bypasses both
the escarpment and the steeply dipping (transgressive
systems tract or TST) slopes, because the slope angles
(35 458) are well above the angle of repose for this size
fraction of material. This fine-grained sediment is subsequently deposited downslope as a wedge of sediment
that Grammer and Ginsburg (1992) termed a highstand wedge (Figure 3). This so-called highstand wedge
occurs in the same position stratigraphically and exhibits the same overall geometry, as might be expected from
a lowstand wedge as represented in classic sequencestratigraphic models (e.g., Sarg, 1988), but will have distinctly different reservoir potential. Based on detailed
sampling and characterization of these fine-grained
highstand sediment aprons, equivalent facies in the
subsurface would be expected to have limited reservoir
potential unless they are interbedded with coarser
grained turbidite and/or debris flow deposits.

New Insights into the Sequencestratigraphic Concept


Seismic imaging has become a standard geophysical
method to image the subsurface architecture of depositional systems and to develop a sequence-stratigraphic
framework. In particular, high-resolution sequence stratigraphy is commonly integrated into the overall workflow for reservoir characterization. In this process, seismic reflections are used as correlation lines, assuming
that they follow depositional surfaces and erosional un-

conformities. This assumption, i.e., that seismic reflections are time lines and thus carry chronostratigraphic
significance, is the basis for the concept of seismic stratigraphy (Vail et al., 1977). Use of seismic horizons as
time lines helps predict ages of strata in undrilled portions of a basin. Because of this predictive capability,
sequence stratigraphy overcomes the shortcomings of
lithostratigraphy, which is commonly time transgressive, and of biostratigraphy and chronostratigraphy,
which are commonly unrelated to physical boundaries
in the rock. Vail et al. (1977) recognized the physical
surfaces on logs and seismic data and postulated that
rapid rates of eustatic change are the major controlling
factor on the timing of stratigraphic discontinuities.
Both of these important assumptions were initially
questioned, but numerous geological and geophysical
experiments have subsequently proved that they are
largely true.
One of these validating experiments was conducted
along the prograding margin of western GBB, with seven
cores in a transect running from the carbonate platform
top to the pelagic sediments in the Straits of Florida
(Eberli et al., 1997; Ginsburg, 2001). Geometries of 17
sequences in the prograding margin suggested that the
pulses of progradation are sea level controlled (Figure 6).
Facies successions in the seven cores corroborate this
seismic interpretation, indicating that global sea level
changes control the architecture of these carbonate sequences (Kievman, 1996; Betzler et al., 1999; Eberli et al.,
2001; Kenter et al., 2001). In particular, facies successions
in the cores contain indications of sea level changes on
two different scales. First, alternating high (as much as
20 cm/k.y.) and low sedimentation rates (<2 cm/k.y.)
record a long-term pattern of bank flooding (0.52 m.y.)
with concomitant shedding to the slope and periods
of bank exposure with reduced shallow-water carbonate production and largely pelagic sedimentation. The
pulses of bank-derived material coincide with prograding pulses that are recognized as seismic sequences.
Second, there are high-frequency alternations between
layers with more platform-derived material separated by
layers with more pelagic sediments. Spectral analyses
document that these alternations are controlled by orbital precession (Williams et al., 2002). Bundling of these
precession-controlled cycles into longer-term obliquity
and eccentricity cycles emerges as a mechanism to create
the lower order, seismically imaged sequences.
Ages of sequence boundaries at the seven sites yield
an excellent correlation between sites, documenting
the age consistency of the sequence boundaries and,
thus, the chronostratigraphic significance of the seismic reflections (Anselmetti et al., 2000; Eberli et al.,
2002). Seismic reflections marking sequence boundaries were dated by means of biostratigraphy in the five
deep-water sites and by a combination of biostratigraphy, magnetostratigraphy, and Sr-isotope stratigraphy

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

FIGURE 6. Line drawing from Western seismic line of the western margin of Great Bahama Bank illustrating the
progradation of the bank and the expansion of the platform facies across the former slope. In the Santaren Channel,
a drift deposit accumulates during the progradation. The wells of the Bahamas transect penetrated each environment.
Small letters indicate seismic sequences of the Neogene strata (modified from Anselmetti et al., 2000).
in the two bank-top sites. The seismic reflection horizons can be traced across a variety of facies belts, from
shallow-water carbonates over slope carbonates to drift
deposits in the Straits of Florida. Despite the fact that
the seismic reflections cross several facies belts, their
ages remain remarkably constant. The average offset in
all sites is 0.38 m.y. The age differences are the result of
biostratigraphic sampling frequency, the spacing of
marker species that required extrapolation of ages, and
the resolution of the seismic data. In no cases, however, do time lines cross seismic reflections.
The fact that the seismic reflections correlate
across the depositional facies transitions and boundaries without changing their chronostratigraphic positions implies that the seismic reflections image the depositional surface instead of lithologic facies. It further
implies that an impedance contrast is generated along
the entire platform-margin transect. To create impedance changes along the entire transect, the lithology
and/or diagenetic modification needs to change along
the entire transect. The investigated seismic reflections,
then, are sequence boundaries representing a fall of sea
level that dramatically changed the sedimentation and
diagenesis in this platform-margin setting (Figure 7).
Carbonate platforms are most productive during
periods of high sea level, when they export abundant
aragonite sediment from the flooded platform top to the
slopes below (e.g., Grammer et al., 1993a; Schlager et al.,
1994). During sea level lowstands, platform production
is reduced, and the relative amount of pelagic deposition, consisting of predominantly calcitic marine microfossils, increases significantly. On the upper or more
proximal portions of the slope, these lowstand units become densely cemented before the next pulse of highstand sediment accumulates (Grammer et al., 1993a,b;
Westphal et al., 1999; Malone et al., 2001). These alternations of well-cemented and less-cemented intervals
on the platform margin and uppermost slope produce

velocity contrasts. In the distal slope and basinal environments, lowstand deposits are generally diagenetically less altered and have lower sonic velocities (Frank
and Bernet, 2000). In addition, a small admixture of
siliciclastic material commonly occurs in lowstand deposits on the slope and basin that further reduces
sonic velocities in the lowstand packages (Isern and
Anselmetti, 2001).
In summary, variability of facies and diagenesis that
occurs in all facies belts along the transect (Figure 7)
are responsible for the acoustic differences and the impedance contrasts that cause seismic reflections along
depositional surfaces, thereby giving them a chronostratigraphic significance. The driving mechanism behind these changes is a fluctuating sea level that causes
changes in sediment composition and diagenetic overprint. Both parameters influence the petrophysical behavior of the strata and, therefore, can cause surfacerelated impedance contrasts.
Vail et al. (1977) cautioned that seismic resolution,
interference, and multiples can produce reflections without time significance. These problems are of increased
importance on the reservoir scale, when seismic resolution is typically pushed to its limit. For example, the
expansion over time of marginal facies over slope facies
along a prograding carbonate platform can produce
strong impedance contrasts and reflections that follow
facies boundaries instead of time lines (Stafleu et al.,
1994; Stafleu and Sonnenfeld, 1994). In some cases, selective filtering of the high-frequency component from
the low-frequency component of the seismic data increases the resolution, so that time-stratigraphic boundaries become visible even if the data is dominated by
lower-frequency components (Zeng and Kerans, 2003).
Seismic modeling of outcrop analogs has proved to be a
powerful method to assess the uncertainty stemming
from resolution problems that might create pseudounconformities (Stafleu and Schlager, 1995).

11

12

Grammer et al.

FIGURE 7. Schematic display of changes in facies, diagenesis, and fluid flow along a carbonate platform margin that
causes impedance contrasts along sequence boundaries along the Bahamas transect. The juxtaposition of sediments
with slightly different composition and diagenetic potential results in density and velocity variations that create the
necessary impedance in each facies belt (modified from Eberli et al., 2002).

Petrophysical Properties of Carbonate


Sediments and Rocks
A major task in reservoir characterization and modeling is to translate geological information into petrophysical properties that can be extracted from geophysical data sets and/or used to populate sedimentary bodies
in reservoir modeling. This task is particularly challenging in carbonates, where cementation and dissolution
processes continuously modify the mineralogy and pore
structure. In extreme cases, this modification can completely reverse the original pore distribution, so that
grains are dissolved to produce pores, whereas the original pore space is filled with cement to form the rock.
All these modifications alter the physical properties of
the rock, thereby resulting in a dynamic relationship
between depositional facies and diagenesis which is
recorded by physical parameters such as porosity, permeability, and sonic velocity. Cores through the modern sediments and Neogene portion of the western
margin of GBB lend themselves to the evaluation of
these changes as they occur in the modern sediment,
then in young rocks, and, ultimately, to completely
cemented or dolomitized rocks.

Modern carbonates display a high spatial homogeneity of petrophysical properties (porosity, density,
and permeability) in each depositional environment
with a good correlation between mechanical energy
and primary characteristics of the sediment, such as
mean grain size and sorting (Incze, 1998). In modern
carbonate sediments, there is typically a negative correlation between porosity and permeability, i.e., the
finest grained sediments have the highest porosity and
the lowest permeability (Enos and Sawatsky, 1981).
This inverse trend is reversed in lithified carbonates,
where the coarse-grained rocks generally have the highest porosity, indicating that diagenesis alters the physical
properties significantly during diagenesis. One of the
first diagenetic processes is compaction, an especially
important process in siliciclastic sediments, which reduces porosity and permeability and thus increases velocity. In carbonates, however, other diagenetic changes
can occur more quickly than compaction, with the result that knowing the compactional history, burial
depth, and/or age is less important for predicting the
physical properties of the rock than in siliciclastics.
In carbonates, cementation and dissolution commonly occur before and simultaneously with compaction.

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

contrast, in a core from the slope of


GBB that was retrieved in a water
depth of 481 m (Ocean Drilling Program Site 1003), diagenesis is not as
rapid, and compaction is better recognized. In this hole, the Vp velocities initially increase with depth in
the unlithified portion of the core.
With the onset of cementation, below
100 m, velocities start to spike in thin
horizons, and at approximately 400
m below the sea floor, velocity values
range from 2000 to 4500 m/s. The
large variability in velocity values indicates that compaction is no longer
the dominant factor for velocity increase and that cementation and dissolution processes thus determine the
elastic behavior of the rock.

Application to Reservoir
Characterization
Seismic data has proved to be
increasingly important in reservoir
characterization. High-resolution 3-D
seismic surveys produce data sets
from which amplitude variations can
be used to interpolate between wells.
FIGURE 8. Illustration of the limited influence of compaction on sonic veReservoir saturation is evaluated using
locity in carbonates. There is a general increase of the lowest velocities with
amplitude variation with offset (AVO),
depth (stippled line); however, large velocity fluctuations at similar depths
and time-lapse surveys delineate prooccur. In Hole Unda, high velocities are present even in the shallow subsurface and are caused by early cementation that is faster than compaction. duction histories and assist in secIn addition, frequent velocity reversals with depth occur in both holes.
ondary recovery. Inversions of seismic volumes into a porosity volume
can be used to predict high-porosity
Marine cementation at depths of 60 m or more can parintervals. Because of the degree of uncertainty in these
tially lithify and transform carbonate sediment across a
geophysical data, accurate interpretation is dependent
platform at rates that are virtually instantaneous geologon the understanding of the rock physics in the imaged
ically (Grammer et al., 1993b; Grammer et al., 1999).
sediments (Mavko et al., 1998). Although sonic velocity
These early marine cements, coupled with shallow subis largely controlled by porosity, many factors, such as
surface processes of cementation, decrease porosity and
clay content and mineralogy, may complicate the readd stiffness to young rocks. The result is that velocity
lationship. This is especially true in carbonates where,
increases much more than would be seen in sediments
as discussed above, velocity is controlled by the comin response solely to compaction.
bined effect of depositional lithology and several postFigure 8 illustrates this phenomenon with logs and
depositional processes that cause a unique velocity disvelocity measurements on discrete samples from two
tribution (Rafavich et al., 1984; Anselmetti and Eberli,
holes located on modern GBB. Although there is a gen1993; Wang, 1997).
eral trend of increasing velocity with depth, velocity
Laboratory measurements from cores in the Bahainversions with depth are common, and values of commas display the porosity-velocity relationship in carbonpressional wave velocities can range as much as 4000 m/s
ates and, together with diagenetic studies, help explain
at approximately the same depth. Furthermore, Vp
the wide scattering of velocity data (Anselmetti and
velocities of more than 4000 m/s are found in shallow
Eberli, 1993, 1997, 2001). Velocity is strongly depenburial depths of less than 20 m at the shallow-water site,
dent on the rock porosity (Wang, 1997; Rafavich et al.,
Unda. At this location, frequent exposure and early ma1984). A plot of porosity vs. velocity displays a clear inrine cementation lithify the sediment prior to burial. In
verse trend; an increase in porosity produces a decrease

13

14

Grammer et al.

moldic rocks and rocks with


interparticle porosity. Moldic
rocks at 4050% porosity can
have Vp as much as 5000 m/s,
whereas rocks with similar
amounts of interparticle porosity or microporosity have
velocities that can be lower
by more than 2500 m/s.
The complicated relationship between porosity and velocity that is observed, which
would also result in a similar
porosity-impedance pattern,
implies that impedance contrasts between two layers can
occur even without a porosity
change, i.e., solely as a result
of different pore types and
pore system architecture. To
further complicate interpreFIGURE 9. Crossplot of Vp (compressional velocity) and porosity of pure carbonate
tation, two layers with differsamples compared to the time average and Woods equation. Velocities of compacted ent porosity values can have
mud samples are only slightly higher than the Woods equation, indicating that these very similar velocities and
samples have nearly no rigidity and that compaction has a minor effect on Vp. The
may have, therefore, no imdiagenetically altered carbonate samples show an inverse relationship between porosity pedance contrast between
and velocity, but there is a large scattering of velocity values at equal porosities and
them. As a result, the scata large range of porosity at a given velocity. This scattering introduces uncertainty
tering in a porosity-velocity
in seismic inversions.
diagram has negative implications for seismic inversion
in velocity (Figure 9). The measured values, however,
and AVO analyses in carbonates. The scattering produces
display a large scatter around this inverse correlation in
an uncertainty in seismic inversion that most current
the velocity-porosity diagram. Velocity differences at
inversion techniques are not able to reduce. For examequal porosities can be more than 2500 m/s, particuple, if a single line from a theoretical equation or a best
larly at higher porosities. For example, rocks with pofit line through the data set is used for inversion, all the
rosities of 39% can have velocities between 2400 and
velocity values above the line will underestimate
5000 m/s. Even at porosities of less than 10%, the veporosity and reserves, whereas all the data points below
locity can still range about 2000 m/s, which is an extrawill overestimate porosity and reserves. Similarly, variaordinary range for rocks with the same chemical comtions in pore type can cause variations in the amplitude
position and the same amount of porosity. Likewise,
with offset that might be more pronounced than varporosity can vary widely at any given velocity. For exiations in saturation or bed thickness. To reduce the
ample, rocks with a Vp of 4100 m/s can have porosities
uncertainties in seismic inversion and AVO analysis,
anywhere between 12 and 43% (Figure 9).
additional study and development of new theoretical
The poor relationship between porosity and velocapproaches are needed that show the physical relationity in carbonates results from the ability of carbonates
ship between pore types, the rock-frame flexibility, and
to form cements and special fabrics with pore types
the elastic behavior in carbonates.
that can enhance the elastic properties of the rock without filling all the pore space. The importance of the pore
type on the elastic property and, thus, the velocity is
OVERVIEW OF PAPERS
illustrated in Figure 10, which shows that different
IN THIS VOLUME
pore types form clusters in the velocity-porosity diagram. The resulting characteristic pattern observed for
The remaining papers of this volume are grouped
every group with the same dominant pore type can
into two broad categories: Reservoir Analogs and
explain why rocks with equal porosity can have very
Integrating Analog Data into Reservoir Modeling.
different velocities. The most prominent velocity contrasts at equal porosities are measured between coarse
The reader will note that the individual papers are

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

FIGURE 10. (A) Graph


of velocity (at 8 MPa
effective pressure) vs.
porosity of various pore
types of carbonates with
an exponential best fit
curve through the data
for reference. Different
pore types cluster in the
porosity-velocity field,
indicating that scattering at equal porosity is
caused by the specific
pore type and its resultant elastic property.
(B) Two pore types with
different elastic behavior. In moldic rocks, pores
are embedded in a stiff
frame, which provides
rigidity to the rock and
results in high velocity
compared to their porosity. In contrast, fabricdestructive dolomitization produces a rock
with intercrystalline porosity with low rigidity
and low velocity (modified from Anselmetti
and Eberli, 1997).

broadly categorized more by the stage the study represents in the progression toward an integrated reservoir model than by the type of reservoir. A brief
summary of the highlights from individual papers
follows.
Chapter 2. Depositional themes of mixed carbonatesiliciclastics in the South Florida Neogene: Application
to ancient deposits, by Donald F. McNeill, Kevin J.
Cunningham, Laura A. Guertin, and Flavio S. Anselmetti.
This paper presents results of a recent drilling project to evaluate the Neogene stratigraphy of south Florida and provides additional insight into the depositional controls and facies patterns of a heterogeneous,

mixed carbonate-siliciclastic depositional system. The


authors present discussion surrounding six key depositional themes from the Neogene and how each theme
may relate to ancient systems: (1) mechanisms by which
antecedent topography controls depositional geometry
and location of depocenters for both carbonates and
siliciclastics; (2) factors that may influence long-distance
transport and mixing of coarse siliciclastics into predominantly carbonate depositional systems; (3) timing
of carbonate platform demise relative to large-scale siliciclastic input; (4) potential for carbonates and siliciclastics in a mixed system to interfinger, forming distinct stratigraphic traps; (5) importance of incorporating

15

16

Grammer et al.

biostratigraphic and chemostratigraphic markers for determining sequence boundaries in some mixed systems
where no clear-cut disconformity exists; and (6) how
shallow burial and early diagenetic processes have produced almost identical acoustic signatures for the two
admixed sediment types.
Chapter 3. Predicting tidal-sand reservoir architecture using data from modern and ancient depositional systems, by Lesli Wood.
Tidally influenced shoreline and deltaic deposits
form some of the largest and most architecturally complicated hydrocarbon fields in the world, but few welldocumented ancient subsurface examples exist. This
paper synthesizes a large data set from the literature on
the dimensions and distribution of modern and ancient siliciclastic tidal sands and examines the utility of
that data set for predicting trends in sand-body dimensions, orientation, and distribution. The paper also presents a comprehensive data set on the dimensions of tidal
sand bars and tidal sand ridges from the Cretaceous Sego
Sandstone in the Book Cliffs of Utah.
Chapter 4. Sequence-stratigraphic and paleogeographic distribution of reservoir-quality dolomite, Madison Formation, Wyoming and Montana, by Langhorne
B. Smith, Jr., Gregor P. Eberli, and Mark Sonnenfeld.
The Mississippian Madison Formation is a classic example of carbonates deposited on a regionally extensive
ramp. The Madison comprises a single, unconformitybounded second-order supersequence that consists internally of a fourfold hierarchy of sequences and cycles.
This paper reviews the sequence-stratigraphic and paleogeographic distribution of porous dolomite in the Madison Formation of Wyoming and Montana and discusses
the implications for exploration-scale variability by
looking at the distribution of dolomite in the sequencestratigraphic framework.
Chapter 5. A laterally accreting grainstone margin from the Albian of northern Mexico: Outcrop model
for Cretaceous reservoirs, by David A. Osleger, Roger
Barnaby, and Charles Kerans.
Outcrop study of progradational, platform-margin
carbonate shoal deposits of Cretaceous age (Albian) in
northern Mexico provides insight into the architecture
of progradational shoal margins and the physical processes that operate along laterally accreting margins
surrounding muddy, intrashelf basins. Results of this
study are compared with subsurface reservoirs in the
Aptian Shuaiba Formation (Bu Hasa field) and the Cenomanian Mishrif Formation of the Arabian Gulf.
Chapter 6. An upper Mississippian carbonate ramp
system from the Pedrogosa Basin, southwestern New
Mexico, United States: An outcrop analog for middle
Carboniferous carbonate reservoirs, by David Sivils.
The Carboniferous (upper Mississippian) Paradise
Formation in the Pedregosa Basin is an outcrop analog
for subsurface reservoirs of a similar age useful for mod-

eling depositional systems, patterns of cyclic sedimentation, reservoir geometry, and diagenetic controls on
reservoir development. This paper focuses on a detailed
outcrop study of the Paradise Formation and how the
results of this study can be used to develop exploration
and exploitation analogs in a sequence-stratigraphic
framework for application to exploration and exploitation of age-equivalent carbonate ramp reservoirs.
Chapter 7. Sedimentology, statistics, and flow behavior for a tide-influenced deltaic sandstone, Frontier
Formation, Wyoming, United States, by Christopher
D. White, Brian J. Willis, Shirley P. Dutton, Janok P.
Bhattacharyra, and Keshav Narayanan.
This paper describes an outcrop study of a
Cretaceous-aged, tidally influenced sandstone exposed
in central Wyoming. The Frewens Allomember of the
Frontier Formation was deposited by a delta prograding into a narrow shoreline embayment between an
older wave-dominated delta lobe and a basin-floor
ridge that was created by subtle structural uplift and
consists of two 5-km-wide by 20-km-long coarseningupward sandstone bodies. Discussion includes details
of the spatial distribution of reservoir facies in outcrop, as well as methods to quantify geologic variability and to predict heterogeneities in the subsurface. Also included is a discussion focusing on the
development of flow models that integrate bedding
geometry, lithofacies, and petrophysical properties in
an appropriate structure for reservoir modeling. These
models were then used to analyze sensitivity of reservoir behavior to different geologic features and to
investigate methods for modeling and upscaling interwell-scale heterogeneity.
Chapter 8. A comparison of two early Miocene
carbonate margins: The Zhujiang carbonate platform
(subsurface, South China Sea) and the Pirinc platform
(outcrop, southern Turkey), by Phil Bassant, Frans Van
Buchem, Andre Strasser, and Anthony Lomando.
The early Miocene Pirinc platform in southern Turkey is an early Miocene carbonate platform-to-basin
transition exposed along the northern flank of the Mut
Basin. The interval formed in response to three highamplitude sea level changes of 100 150-m amplitude
that resulted in a steep-edged platform with prograding slope and margin geometries during highstands,
with thin carbonate platforms characterized by onlapping geometries during lowstands. Results of the outcrop study in Turkey are compared with the subsurface
early Miocene Zhujiang platform in China and lead to
exploration- and production-scale play concepts that
may be particularly appropriate to early Miocene carbonate reservoirs in Southeast Asia.
Chapter 9. Accommodation-controlled systems
tract-specific facies partitioning and resulting geometric development of reservoir grainstone ramp-crest shoal
bodies, by Victoria L. French and Charles Kerans.

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

High-resolution sequence-stratigraphic studies from


Permian outcrop analogs in the Guadalupe Mountains
of west Texas and New Mexico, and from the subsurface
West Jordan San Andres unit (Permian) in Texas, show
that both stochastic and deterministic reservoir models
can be greatly improved by accounting for both systems tract specific facies partitioning and incorporation of outcrop-based object data from comparable
stratigraphic/accommodation settings. Both the outcrop and subsurface models exhibit a high degree of
facies-dependent compartmentalization that is a result
of varying accommodation. In addition to documenting the degree of vertical and lateral heterogeneity that
should be expected in reservoirs producing from highstand carbonate shoal complexes, the study also provides a predictive geometric relationship between shoalbody maximum thickness and expected maximum dip
dimensions.
Chapter 10. Reservoir characterization in the San
Andres Formation of Vacuum field, Lea County, New
Mexico: Another use of the San Andres Algerita outcrop model for improved reservoir description, by Emily
L. Stoudt and Michael A. Raines.
Early geologic models of Vacuum field, New Mexico
(Permian, San Andres Formation reservoir) were generated primarily from petrophysical data and resulted in
lithostratigraphic correlations that crossed time lines
and flow unit boundaries. The application of a chronostratigraphic framework previously developed in outcrop led to an improved reservoir model for Vacuum
field. The improved model includes (1) the recognition
of compartmentalized tidal flat cycles; (2) the identification of bypassed pay in porous dolopackstones
that are stratigraphically equivalent to, but downdip
from, the tidal flats; and (3) the recognition of the presence of tight karst intervals in older high-frequency
sequences that compartmentalize the most continuous
San Andres pay interval.
Chapter 11. An integrated approach to characterization and modeling of deep-water reservoirs,
Diana field, western Gulf of Mexico, by Morgan Sullivan, Lincoln Foreman, David Jennette, David Stern,
Gerrick Jensen, and Frank Goulding.
Deep-water outcrop analog data from the Lower
Permian Skoorsteenberg Formation in the Tanqua Karoo
Basin of South Africa and the Upper Carboniferous Ross
Formation in the Clare Basin, western Ireland, were
integrated with seismic and well data to better characterize the Diana field in the western Gulf of Mexico.
The Diana field is typical of many fields in the early
stages of development, with a reservoir having limited
seismic resolution and subsurface data sets limited by
well spacing. Using bed-scale reservoir architectures
that were quantified with photomosaics and by correlation of closely spaced measured sections, a spectrum of channel dimensions and shapes were compiled

from the outcrop to facilitate the conditioning of a 3-D


object-based geologic model for the Diana field. Advantages of the resulting Diana field model are that (1)
it incorporates geologic interpretations, (2) it honors
all available data, and (3) it models the reservoir as
discrete objects with specific dimensions, facies juxtapositions, and connectivity. The result is a study that
provides the framework for optimal placement of wells
to maximize architectural and facies controls on reservoir performance.
Chapter 12. Outcrop-based 3-D modeling of the
Tensleep Sandstone at Alkali Creek, Bighorn Basin,
Wyoming, by Bozkurt N. Ciftci, A. A. Aviantara, Neil F.
Hurley, and D. R. Kerr.
To identify the geometry and volumetric size of
eolian compartments, defined here as a body of
rock that is surrounded by eolian bounding surfaces, a
3-D computer model of the Carboniferous to Permian
Tensleep Sandstone was generated from outcrop data in
the Bighorn Basin of Wyoming. The bounding-surface
hierarchy was mapped by traditional surveying methods and a precise Global Positioning System (GPS) receiver to establish the input data for the modeling procedures. The model was built from correlative bounding
surfaces observed in the walls of parallel canyons that
cut down into the Tensleep Sandstone. Wells were simulated by 10-, 20-, 40-, 80-, and 160-ac templates in the
3-D model and included simulation of horizontal wells
in different orientations (parallel, perpendicular, and
oblique to foreset dip direction). Results from the modeling indicate that horizontal wells drilled parallel to
foreset dip direction drain the maximum number and
volume of reservoir compartments.
Chapter 13. Integration of high-resolution outcrop and subsurface data to enhance interpretation of
low-resolution seismic data in the Upper Devonian
(Frasnian) carbonate system in western Canada, by
Anne M. Schwab, Frans S. P. Van Buchem, and Gregor
P. Eberli.
The Upper Devonian Miette buildup of the Canadian Rockies provides a representative example of the
sequence-stratigraphic framework of the Late Devonian
for the subsurface of Alberta, Canada. A synthetic seismic section across the Miette buildup in outcrop is used
for calibrating and comparing the outcrop geometries
and facies with seismic and log data across the subsurface Redwater reef complex. The iterative interpretation of seismic data and well logs with the outcrop data
results in a reinterpretation of the platform architecture
of the Redwater reef complex. This reinterpretation is
based on (1) the identification of four sedimentological
phases in platform construction that form in a secondorder sea level cycle, and (2) details of six third-order
sequences that can be correlated to both facies in core,
and the geometries observed in the seismic data from
the Redwater reef complex. Because this second-order

17

18

Grammer et al.

sea level change in the Frasnian is mostly eustatic in


nature, the resulting platform architecture is expected
to be found in other age-equivalent isolated carbonate
buildups in Canada and elsewhere in the world.
Chapter 14. Well placement, cost reduction, and
increased production using reservoir models based on
outcrop, core, well logs, seismic data, and modern analogs: Onshore and offshore western Trinidad, by Grant
D. Wach, Chris S. Lolley, Donald S. Mims, and Clyde A.
Sellers.
Fluvial/estuarine complexes are significant producing reservoirs both onshore and offshore western Trinidad, but the channel complexes are very difficult to
correlate in the subsurface. Numerous permeability baffles and barriers create complex reservoir heterogeneity
that can result in significant bypassed hydrocarbons.
Outcrops of tidally influenced nonmarine channel complexes and modern depositional analogs are used to
determine the architectural elements and bounding
surfaces that impact reservoir continuity and heterogeneity, thereby highlighting possible subsurface correlation problems. Two reservoir models are developed
and discussed. The first is of a nonmarine channel complex based mainly on outcrop data, where two reservoir
simulations help characterize the architectural elements
and heterogeneity of the tidally influenced channel
system. The second model incorporates outcrop data
and modern depositional analogs to examine channel
systems in the Soldado field, offshore western Trinidad,
identified by examination of cores, well logs, and seismic data.
Chapter 15. Integrating sequence stratigraphy and
multiple 3-D geostatistical realizations in constructing
a model of the second Eocene reservoir, Wafra field,
Partitioned Neutral Zone, Kuwait and Saudi Arabia, by
Dennis W. Dull.
Sequence-stratigraphic principals and 3-D geostatistical modeling were integrated to create a geologic
model of the second Eocene reservoir in the Wafra field.
The reservoir consists of four high-frequency sequences
stacked into an overall regressive pattern in a carbonate
ramp and characterized by complex dolomitized reservoirs. The modeling of the second Eocene was accomplished with the following three steps: (1) establishing
a relationship between core-derived lithofacies and log
response, (2) constructing the sequence-stratigraphic
framework, and (3) applying geostatistics to integrate
the lithofacies, sequence-stratigraphic framework, and
spatial correlation. Construction of the reservoir model in an integrated sequence-stratigraphic framework
resulted in the identification of new potential reservoir
intervals in the field, as well as defining a large number
of infill drilling locations.
Chapter 16. Outcrop and waterflood simulation
modeling of the 100-Foot Channel Complex, Texas,
and the Ainsa II channel complex, Spain: Analogs to

multistory and multilateral channelized slope reservoirs, by David K. Larue.


Two outcrop examples of deep-water depositional
systems are discussed, the 100-Foot Channel Complex in the Permian Brushy Canyon Formation of
southwest Texas, and the Ainsa II channel complex in
the Eocene Hecho Group of Spain. Conceptual geologic
data from other outcrop studies, as well as 3-D seismic
studies, were incorporated into detailed 2-D outcrop
realizations to construct 3-D reservoir models. The two
outcrops studied are significant in that the 100-Foot
Channel Complex is a multistory channel complex,
whereas the Ainsa II channel complex is both multistory and multilateral in distribution. Because of the
conceptual nature of the study, multiple models were
created to account for uncertainty 4 models for the
100-Foot Channel Complex and 11 models for the
Ainsa II channel complex. Models for the 100-Foot
Channel Complex addressed a question regarding the
importance of siltstone continuity in channel fills in
this multistoried channel-fill complex. According to
the models for the Ainsa channel complex, permeability, heterogeneity and sandstone connectivity appear to
be the most significant factors influencing waterflood
recovery.
Chapter 17. Computer simulation of carbonate
sedimentary and shallow diagenetic processes, by Fumiaki Matsuda, Michinori Saito, Ryotaro Iwahashi, Hiroshi Oda, and Yoshihiro Tsuji.
A newly developed computer simulation model,
Facies-3D, designed to model both sedimentary and
early diagenetic processes in both carbonates and siliciclastics, is used to model the 3-D distribution of carbonate facies based on water depth and current velocity
in conjunction with paleotopography. The paper discusses the principles of the Facies-3D model and the
results of two simulation case studies. The simulation
case studies are presented using an established depositional and diagenetic model of carbonate reservoirs
of the Pleistocene Ryuku Group in southwestern Japan
and of the upper Miocene Kais Formation of Irian Jaya,
Indonesia. In the simulation case studies of the two
depositional models, the Facies-3D model provided a
close approximation of the distribution of facies with
respect to water depth, current velocity related to paleotopography, and the degree of meteoric diagenesis in
relationship to paleotopography and sea level change.
Chapter 18. Multiple-point geostatistics: A quantitative vehicle for integrating geologic analogs into
multiple reservoir models, by Jef Caers and Tuanfeng
Zhang.
Traditional variogram-based geostatistics is inadequate at capturing geological heterogeneity from outcrop data because the variogram is too limiting. This
paper discusses a new approach, termed multiplepoint geostatistics, that does not rely on variogram

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

models. Multiple-point geostatistics borrows multiplepoint patterns from what are termed training images
then anchors them to subsurface well-log, seismic, and
production data. The paper outlines the guiding principles of using analog models in multiple-point geostatistics and shows that simple so-called modular training
images can be used to build complex reservoir models
using geostatistical algorithms.

ACKNOWLEDGMENTS
We extend our sincere thanks to the authors of the
papers in this volume for their interesting and timely
contributions to this memoir. Many thanks also to
AAPG, and especially the publications staff, who were
invaluable in resolving many of the logistical issues
faced by both the editors and the authors of the volume. We also extend our thanks to Tony Sandomierski
for assistance with figures and Susan Friel of Biotechwrite for editorial suggestions. Lastly, we would like to
thank the following individuals, whose critical reviews
of the manuscripts assured us of high quality and
comprehensive compilations of current progress in
this rapidly changing field of study: Rick Abegg, Dave
Barnes, Jennifer Beall, Terry Belsher, Jim Borer, Jean
Borgomono, Bryan Bracken, Trevor Burchette, Mary
Carr, Chris Crescini, Tim Cross, Steve Dorobek, Evan
Franseen, Ray Garber, Jean Hsieh, Jim Jennings, Chris
Kendall, Charlie Kerans, Rebecca Latimer, Marge Levy,
Bob Lindsay, Susan Longacre, Jerry Lucia, Sal Mazzullo,
Tim McHargue, Bill Morgan, John Pendrel, Carlos
Pirmez, Mike Pope, Dennis Prezbindowski, Gene
Rankey, Will Schweller, and John Weisenberger.

REFERENCES CITED
Amos, C. L., and E. L. King, 1984, Bedforms of the Canadian eastern seaboard: A comparison with global occurrences: Marine Geology, v. 57, p. 167 208.
Anselmetti, F. S., and G. P. Eberli, 1993, Controls on sonic
velocity in carbonates: Pure and Applied Geophysics,
v. 141/2-4, p. 287 323.
Anselmetti, F. S., and G. P. Eberli, 1997, Sonic velocity in
carbonate sediments and rocks, in F. J. Marfurt and A.
Palaz, eds., Carbonate seismology: Society of Exploration Geophysicists, Geophysical Development Series no. 6, p. 53 74.
Anselmetti, F. S., and G. P. Eberli, 2001, Sonic velocity in
carbonates A combined product of depositional lithology and diagenetic alterations, in R. N. Ginsburg,
ed., Subsurface geology of a prograding carbonate
platform margin, Great Bahama Bank: Results of the
Bahamas Drilling Project: SEPM Special Publication
70, p. 193 216.

Anselmetti, F. S., G. P. Eberli, and Z.-D. Ding, 2000, From


the Great Bahama Bank into the Straits of Florida: A
margin architecture controlled by sea level fluctuations and ocean currents: Geological Society of America Bulletin, v. 112, p. 829 846.
Ball, M. M., 1967, Carbonate sand bodies of Florida and
the Bahamas: Journal of Sedimentary Petrology, v. 37,
p. 556 591.
Beach, D. K., 1993, Submarine cementation of subsurface
Pliocene carbonates from the interior of Great Bahama Bank: Journal of Sedimentary Petrology, v. 63,
p. 1059 1069.
Beach, D. K., and R. N. Ginsburg, 1980, Facies succession,
Plio-Pleistocene carbonates, northwestern Great Bahamas Bank: AAPG Bulletin, v. 64, p. 1634 1642.
Betzler, C., J. J. G. Reijmer, K. Bernet, G. P. Eberli, and F. S.
Anselmetti, 1999, Sedimentary patterns and geometries of the Bahamian outer carbonate ramp (Miocene
and lower Pliocene, Great Bahama Bank): Sedimentology, v. 46, p. 1127 1145.
Borer, J. M., and P. M. Harris, 1991, Depositional facies
and cyclicity in the Yates Formation, Permian Basin
Implications for reservoir heterogeneity: AAPG Bulletin, v. 75, p. 726 779.
Camacho, H., C. J. Busby, and B. Kneller, 2002, A new
depositional model for the classical turbidite locality
at San Clemente State Beach, California: AAPG Bulletin, v. 86, p. 1543 1560.
Cook, H. E., 1983, Ancient carbonate platform margins,
slopes, and basins: Platform margin and deep water
carbonates: SEPM Short Course no. 12, p. 5.1 5.189.
Cook, H. E., and H. T. Mullins, 1983, Basin margin, in P. A.
Scholle, D. G. Bebout, and C. H. Moore, eds., Carbonate depositional environments: AAPG Memoir 33,
p. 539 619.
Davis, R. A., and P. S. Balson, 1992, Stratigraphy of a
North Sea tidal sand ridge: Journal of Sedimentary
Petrology, v. 62, p. 116 121.
Dawans, J. M., and P. K. Swart, 1988, Textural and
geochemical alterations in late Cenozoic Bahamian
dolomites: Sedimentology, v. 35, p. 385 403.
Dulaney, J. P., and A. L. Hadik, 1990, Geologic reservoir
description of Mobil-operated units in Slaughter (San
Andres) field, Cochran and Hockley Counties, Texas,
in D. G. Bebout and P. M. Harris, eds., Geologic and
engineering approaches in evaluation of San Andres/
Grayburg hydrocarbon reservoirs Permian Basin:
Bureau of Economic Geology, University of Texas at
Austin, p. 53 73.
Dyer, K. A., and D. A. Huntley, 1999, The origin, classification and modeling of sand banks and ridges: Continental Shelf Research, v. 19, p. 1285 1330.
Eberli, G. P., and R. N. Ginsburg, 1987, Segmentation
and coalescence of Cenozoic carbonate platforms,
northwestern Great Bahama Bank: Geology, v. 15,
p. 75 79.
Eberli, G. P., P. K. Swart, M. Malone, and Scientific Party,
1997, Proceedings of the Ocean Drilling Program,
Initial Reports, 166: College Station, Texas, Ocean
Drilling Program, 850 p.
Eberli, G. P., F. S. Anselmetti, J. A. M. Kenter, D. F. McNeill,

19

20

Grammer et al.

and L. A. Melim, 2001, Calibration of seismic sequence stratigraphy with cores and logs, in R. N.
Ginsburg, ed., Subsurface geology of a prograding carbonate platform margin, Great Bahama Bank: SEPM
Special Publication 70, p. 241 265.
Eberli, G. P., F. S. Anselmetti, D. Kroon, T. Sato, and J. D.
Wright, 2002, The chronostratigraphic significance
of seismic reflections along the Bahamas Transect:
Marine Geology, v. 185, nos. 1 2, p. 1 17.
Enos, P., and L. H. Sawatsky, 1981, Pore networks in
Holocene carbonate sediments: Journal of Sedimentary Petrology, v. 51, p. 961 985.
Enos, P., and C. H. Moore, 1983, Fore-reef slope, in P. A.
Scholle, D. G. Bebout, and C. H. Moore, eds., Carbonate depositional environments: AAPG Memoir
33, p. 507 539.
Frank, T. D., and K. Bernet, 2000, Isotopic signature of
burial diagenesis and primary lithologic contrasts in
periplatform carbonates (Miocene, Great Bahama
Bank): Sedimentology, v. 47, p. 1256 1267.
Ginsburg, R. N., ed., 2001, Subsurface geology of a
prograding carbonate platform margin, Great
Bahama Bank: Results of the Bahamas Drilling
Project: SEPM Special Publication 70, 265 p.
Ginsburg, R. N., and N. P. James, 1974, Holocene carbonate sediments of continental shelves, in C. A.
Burk and C. L. Drake, eds., The geology of continental margins: New York, Springer-Verlag, p. 137
155.
Ginsburg, R. N., P. M. Harris, G. P. Eberli, and P. K. Swart,
1991, The growth potential of a bypass margin, Great
Bahama Bank: Journal of Sedimentary Petrology,
v. 61, no. 1, p. 976 987.
Grammer, G. M., 1991, Formation and evolution of Quaternary carbonate foreslopes, Tongue of the Ocean, Bahamas: Unpublished Ph.D. dissertation, University of
Miami, Rosenstiel School of Marine and Atmospheric
Science, Coral Gables, Florida, 375 p.
Grammer, G. M., and R. N. Ginsburg, 1992, Highstand vs.
lowstand deposition on carbonate platform margins:
Insight from Quaternary foreslopes in the Bahamas:
Marine Geology, v. 103, p. 125 136.
Grammer, G. M., R. N. Ginsburg, and P. M. Harris, 1993a,
Timing of deposition, diagenesis, and failure of steep
carbonate slopes in response to a high-amplitude/
high-frequency fluctuation in sea level, Tongue of the
Ocean, Bahamas, in R. G. Loucks and J. F. Sarg, eds.,
Carbonate sequence stratigraphy, recent developments
and applications: AAPG Memoir 57, p. 107 131.
Grammer, G. M., R. N. Ginsburg, P. K. Swart, D. F. McNeill,
A. J. T. Jull, and D. R. Prezbindowski, 1993b, Rapid
growth rate of syndepositional marine aragonite cements in steep marginal slope deposits, Bahamas and
Belize: Journal of Sedimentary Petrology, v. 63, p. 983
989.
Grammer, G. M., G. P. Eberli, F. S. P. Van Buchem, G. M.
Stevenson, and P. Homewood, 1996, Application of
high-resolution sequence stratigraphy to evaluate
lateral variability in outcrop and subsurface Desert
Creek and Ismay intervals (Pennsylvanian), Paradox
Basin, in M. Sonnenfeld and M. Longman, eds.,

Paleozoic systems of the Rocky Mountain region:


Rocky Mountain Section SEPM Special Publication,
p. 235 266.
Grammer, G. M., C. M. Crescini, D. F. McNeill, and
L. H. Taylor, 1999, Quantifying rates of syndepositional marine cementation in deeper platform
environmentsNew insight into a fundamental process: Journal of Sedimentary Research, v. 69, p. 202
207.
Grammer, G. M., P. M. Harris, and G. P. Eberli, 2001,
Carbonate platforms: Exploration and production
scale insight from modern analogs in the Bahamas:
The Leading Edge, Society of Economic Geophysicists, v. 20, p. 252 261.
Grant, C. W., D. J. Goggin, and P. M. Harris, 1994, Outcrop analog for cyclic-shelf reservoirs, San Andres
Formation of Permian Basin: Stratigraphic framework, permeability distribution, geostatistics, and
fluid-flow modeling: AAPG Bulletin, v. 78, p. 23 54.
Halley, R. B., and P. M. Harris, 1979, Fresh-water cementation of a 1000-year-old oolite: Journal of Sedimentary Petrology, v. 49, p. 969 987.
Halley, R. B., P. M. Harris, and A. C. Hine, 1983, Bank
margin environments, in P. A. Scholle, D. G. Bebout,
and C. H. Moore, eds., Carbonate depositional environments: AAPG Memoir 33, p. 463 506.
Hardie, L. A., 1977, ed., Sedimentation on the modern
carbonate tidal flats of northwest Andros Island,
Bahamas: Johns Hopkins University, Studies in
Geology, no. 22, 202 p.
Harris, P. M., 1978, Holocene marine-cemented sands,
Joulters ooid shoal, Bahamas: Transactions of the
Gulf Coast Association of Geological Societies, v. 28,
p. 175 183.
Harris, P. M., 1979, Facies anatomy and diagenesis of a
Bahamian ooid shoal: Sedimenta 7: Miami Beach,
Florida, Comparative Sedimentology Laboratory,
University of Miami, 163 p.
Harris, P. M., 1983, The Joulters ooid shoal, Great Bahama Bank, in T. Peryt, ed., Coated grains: New York,
Springer-Verlag, p. 132 141.
Harris, P. M., 1984, Cores from a modern sand body; the
Joulters ooid shoal, Great Bahama Bank, in P. M.
Harris, ed., Carbonate sands A core workshop:
SEPM Core Workshop no. 5, p. 429 464.
Harris, P. M., and S. D. Walker, 1990, McElroy field:
Development geology of a dolostone reservoir, Permian Basin, west Texas, in D. G. Bebout and P. M.
Harris, eds., Geologic and engineering approaches in
evaluation of San Andres/Grayburg hydrocarbon
reservoirs Permian Basin: Bureau of Economic Geology, University of Texas at Austin, p. 275 296.
Hine, A. C., and A. C. Neumann, 1977, Shallow carbonatebank-margin growth and structure, Little Bahama
Bank, Bahamas: AAPG Bulletin, v. 61, p. 376 406.
Hine, A. C., R. J. Wilber, and A. C. Neumann, 1981,
Carbonate sand bodies along contrasting shallow
bank margins facing open seaways in northern
Bahamas: AAPG Bulletin, v. 65, p. 261 290.
Hulscher, S. J. M. H., H. E. De Swart, and H. J. De Vriend,
1993, The generation of offshore tidal sand banks

Integration of Outcrop and Modern Analogs in Reservoir Modeling: Overview

and sand waves: Continental Shelf Research, v. 13,


p. 1183 1204.
Huthnance, J. M., 1982, On one mechanism forming linear sand banks: Estuarine and Coastal Marine Science, v. 145, p. 79 99.
Illing, L. V., 1954, Bahaman calcareous sands: AAPG Bulletin, v. 38, p. 1 95.
Incze, M. L., 1998, Petrophysical properties of shallowwater carbonates in modern depositional and shallow sub-surface environments: Unpublished Ph.D.
dissertation, University of Miami, Rosenstiel School
of Marine and Atmospheric Science, Coral Gables,
Florida, 405 p.
Isern, A. R., and F. S. Anselmetti, 2001, The influence of
carbonate platform morphology and sea level on
fifth-order petrophysical cyclicity in slope and basin
sediments adjacent to the Great Bahama Bank:
Marine Geology, v. 177, p. 381 394
Kenter, J. A. M., R. N. Ginsburg, and S. R. Troelstra, 2001,
The western Great Bahama Bank: Sea-level-driven
sedimentation patterns on the slope and margin, in
R. N. Ginsburg, ed., Subsurface geology of a prograding carbonate platform margin, Great Bahama Bank:
SEPM Special Publication 70, p. 61 100.
Kerans, C., F. J. Lucia, and R. K. Senger, 1994, Integrated
characterization of carbonate ramp reservoirs using
outcrop analogs: AAPG Bulletin, v. 78, p. 181 216.
Kerans, C., and S. W. Tinker, 1997, Sequence stratigraphy
and characterization of carbonate reservoirs: SEPM
Short Course Notes, no. 40, 130 p.
Kievman, C. M., 1996, Sea level effects on carbonate
platform evolution: Plio-Pleistocene, northwestern
Great Bahama Bank: Unpublished Ph.D. dissertation, University of Miami, Rosenstiel School of Marine and Atmospheric Science, Coral Gables, Florida,
245 p.
Lucia, F. J., D. G. Bebout, and C. R. Hocott, 1990, Reservoir characterization through integration of geological and engineering methods and techniques: Dune
(Grayburg) field, University Lands, Crane County,
Texas, in D. G. Bebout and P. M. Harris, eds., Geologic and engineering approaches in evaluation of
San Andres/Grayburg hydrocarbon reservoirs Permian Basin: Bureau of Economic Geology, University
of Texas at Austin, p. 197 238.
Malone, M. J., N. C. Slowey, and G. M. Henderson, 2001,
Early diagenesis of shallow-water periplatform carbonate sediments, leeward margin, Great Bahama
Bank (Ocean Drilling Program Leg 166): Geological
Society of America Bulletin, v. 113, p. 881 894.
Major, R. P., G. W. Vander Stoep, and M. H. Holtz, 1990,
Geological and engineering assessment of remaining
mobile oil, East Penwell San Andres Unit, Ector
County, Texas, in D. G. Bebout and P. M. Harris,
eds., Geologic and engineering approaches in evaluation of San Andres/Grayburg hydrocarbon reservoirs Permian Basin: Bureau of Economic Geology,
University of Texas at Austin, p. 175 196.
Major, R. P., D. G. Bebout, and P. M. Harris, 1996, Facies
heterogeneity in a modern ooid sand shoal An
analog for hydrocarbon reservoirs: Bureau of Eco-

nomic Geology, Geological Circular 96-1, University


of Texas at Austin, 30 p.
Mavko, G., T. Mukerji, and J. Dvorkin, 1998, The rock
physics handbook; Tools for seismic analysis in
porous media: Cambridge, U.K., Cambridge University Press, 329 p.
McBride, R. A., and T. F. Moslow, 1991, Origin, evolution
and distribution of shoreface sand ridges, Atlantic
inner shelf, U.S.A.: Marine Geology, v. 97, p. 57 85.
McLaurin, B. T., and R. J. Steele, 2000, Fourth-order nonmarine to marine sequences, Middle Castlegate Formation, Book Cliffs, Utah: Geology, v. 28, p. 359 362.
Melim, L. A., P. K. Swart, and R. G. Maliva, 1995, Meteoriclike fabrics forming in marine waters: Implications
for the use of petrography to identify diagenetic environments: Geology, v. 23, p. 755 758.
Melim, L. A., H. Westphal, P. K. Swart, G. P. Eberli, and A.
Munnecke, 2002, Questioning carbonate diagenetic
paradigms: Evidence from the Neogene of the
Bahamas: Marine Geology, v. 185, p. 27 53.
Newell, N. D., and J. K. Rigby, 1957, Geological studies on
the Great Bahama Bank, in R. J. LeBlanc and J. G.
Breeding, eds., Regional aspects of carbonate deposition: A symposium with discussion: SEPM Special
Publication 5, p. 15 72.
Off, T., 1963, Rhythmic linear sand units caused by tidal
currents: AAPG Bulletin, v. 47, p. 324 341.
Purdy, E. G., 1963, Recent calcium carbonate facies of the
Great Bahama Bank: Journal of Geology, v. 71,
p. 472 497.
Purves, W. J., 1990, Reservoir description of the Mobil Oil
Bridges State Leases (upper San Andres reservoir),
Vacuum field, Lea County, New Mexico, in D. G.
Bebout and P. M. Harris, eds., Geologic and engineering approaches in evaluation of San Andres/
Grayburg hydrocarbon reservoirs Permian Basin:
Bureau of Economic Geology, University of Texas at
Austin, p. 87 112.
Rafavich, F., C. G. St. C. Kendall, and T. P. Todd, 1984,
The relationship between acoustic properties and the
petrographic character of carbonate rocks: Geophysics, v. 49, p. 1622 1636.
Ruppel, S. C., 1990, Facies control of porosity and
permeability: Emma San Andres reservoir, Andrews
County, Texas, in D. G. Bebout and P. M. Harris, eds.,
Geologic and engineering approaches in evaluation
of San Andres/Grayburg hydrocarbon reservoirs
Permian Basin: Bureau of Economic Geology, University of Texas at Austin, p. 145 171.
Sarg, J. F., 1988, Carbonate sequence stratigraphy, 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 level changes An integrated approach: SEPM
Special Publication 42, p. 155 181.
Sarg, J. F., J. R. Markello, and L. J. Weber, 1999, The secondorder cycle, carbonate-platform growth, and reservoir,
source, and trap prediction, in P. M. Harris, A. H. Saller,
and J. A. Simo, eds., Advances in carbonate sequence
stratigraphy: Application to reservoirs, outcrops, and
models: SEPM Special Publication 63, p. 11 34.
Schlager, W., J. J. G. Reijmer, and A. W. Droxler, 1994,

21

22

Grammer et al.

Highstand shedding of carbonate platforms: Journal


of Sedimentary Research, v. B64, p. 270 281.
Shinn, E. A., R. M. Lloyd, and R. N. Ginsburg, 1969,
Anatomy of a modern carbonate tidal-flat, Andros
Island, Bahamas: Journal of Sedimentary Petrology,
v. 39, p. 1202 1228.
Stafleu, J., and M. D. Sonnenfeld, 1994, Seismic models of
a shelf-margin depositional sequence: Upper San
Andres Formation, Last Chance Canyon, New Mexico: Journal of Sedimentary Research, v. B64/4,
p. 481 499.
Stafleu, J., and W. Schlager, 1995, Pseudounconformities
in seismic models of large outcrops: Geologische
Rundschau, v. 84, p. 761 769.
Stafleu, J., A. J. W. Everts, and J. A. M. Kenter, 1994,
Seismic models of a prograding carbonate platform:
Vercors, SE France: Marine and Petroleum Geology,
v. 11, p. 514 517.
Tinker, S. W., 1996, Building the 3-D jigsaw puzzle:
Applications of sequence stratigraphy to 3-D reservoir characterization, Permian Basin: AAPG Bulletin,
v. 80, p. 460 485.
Tyler, N., and R. J. Finley, 1991, Architectural controls on
the recovery of hydrocarbons from sandstone reservoirs, in A. D. Miall and N. Tyler, eds., The threedimensional facies architecture of terrigenous clastic
sediments and its implications for hydrocarbon
discovery and recovery: SEPM Concepts in Sedimentology and Paleontology, v. 3, p. 1 5.
Vail, P. R., A. G. Todd, and J. B. Sangree, 1977, Chronostratigraphic significance of seismic reflections, in
C. E. Payton, ed., Seismic stratigraphy Applications
to hydrocarbon exploration: AAPG Memoir 26,
p. 99 116.
Wang, Z., 1997, Seismic properties of carbonate rocks,
in I. Palaz and K. J. Marfurt, eds., Carbonate seismology: Society of Exploration Geophysicists Geophysical Development Series no. 6, p. 29 52.

Weil, C. B., R. D. Morse, and R. E. Sheridan, 1973, A


model for the evolution of linear tidal-built sand
shoals in Delaware Bay, in Relations sedimentaires
entre estuaries et plateaux continentaux: Bordeaux,
France, Institut de Geologie du Bassin dAquitaine,
p. 108.
Westphal, H., J. J. G. Reijmer, and M. J. Head, 1999,
Sedimentary input and diagenesis on a carbonate
slope (Bahamas): Response to morphology evolution
of the carbonate platform and sea level fluctuations,
in P. M. Harris, A. H. Saller, and A. J. Simo, eds.,
Advances in carbonate sequence stratigraphy Application to reservoirs, outcrops, and models: SEPM
Special Publication 63, p. 247 274.
White, C. D., and M. D. Barton, 1999, Translating
outcrop data to flow models, with applications to
the Ferron Sandstone: Society of Petroleum Engineers
Reservoir Evaluation and Engineering, v. 2, p. 341
350.
Wilber, R. J., J. D. Milliman, and R. B. Halley, 1990,
Accumulation of bank-top sediment on the western
slope of Great Bahama Bank: Rapid progradation of a
carbonate megabank: Geology, v. 18, p. 970 974.
Williams, T., D. Kroon, and S. Spezzaferri, 2002, Middle
upper Miocene cyclostratigraphy of downhole logs
and short to long term astronomical cycles in
carbonate production of Great Bahama Bank: Marine
Geology, v. 185, p. 75 93.
Willis, B. J., and C. D. White, 2000, Quantitative outcrop
data for flow simulation: Journal of Sedimentary
Research, v. 70, p. 788 802.
Willis, B., and S. Gabel, 2001, Sharp-based, tide-dominated deltas of the Sego Sandstone, Book Cliffs, Utah,
U.S.A.: Sedimentology, v. 48, p. 479 506.
Zeng, H., and C. Kerans, 2003, Seismic frequency control
on carbonate seismic stratigraphy: A case study of the
Kingdom Abo sequence, west Texas: AAPG Bulletin,
v. 87, p. 273 293.

Anda mungkin juga menyukai