Sei sulla pagina 1di 265

CONTENTS

PAGE NO

I. SUBSURFACE FACIES ANALYSIS


1. SEDIMENTARY ENVIRONMENTS AND FACIES 3
2. LITHOLOGY 11
3. SEDIMENTARY STRUCTURES, GEOMETRY & FACIES 16
4. PALAEONTOLOGY 26
5. SOURCES OF INFORMATION 35
6. SANDSTONE RESERVOIR FACIES 52
7. CARBONATE RESERVOIR FACIES 95

II. SEQUENCE STRATIGRAPHY


1. GENERAL CONCEPTS 125
2. CYCLOSTRATIGRAPHY 136
3. THE MAKING OF A STRATIGRAPHIC SEQUENCE 140
4. SEQUENCE STRATIGRAPHY 151
5. INTERPRETATION OF STRATIGRAPHIC SEQUENCES 173
6. CONCLUSIONS AND CURRENT CONTROVERSIES 184

III. SANDSTONE RESERVOIRS


1. NON MARINE SANDSTONE RESERVOIRS 193
2. MARGINAL MARINE SANDSTONE RESERVOIRS
3. MARINE SANDSTONE RESERVOIRS
4. DEEP MARINE SANDSTONE RESERVOIRS
I. SUBSURFACE FACIES ANALYSIS

1. SEDIMENTARY ENVIRONMENTS AND FACIES

Sedimentary Environments
The essence of facies analysis is the construction of a geologic model that describes an ancient
sedimentary environment. Consistent with Lyell's doctrine of uniformitarianism -- "the present is a
key to the past "-- modern equivalents are used as analogs of ancient environments. But
explorationists must be cautious in using modern analogs to reconstruct ancient environments.
Basic geological processes involved in erosion, transportation, and deposition have probably not
changed throughout geologic time. In ancient times, though, factors that controlled these
processes (such as climate, tidal currents, and vegetation) were very likely quite different from
today's. Certainly vegetation, which exerts a strong control on erosional processes, was absent or
sparse prior to midPaleozoic times.

According to Selley (1985) a sedimentary environment "is a part of the earth's surface which is
physically, chemically, and biologically distinct from adjacent terraines." A sedimentary
environment may be characterized by erosion, equilibrium, or deposition. Classifications of the
main terrigenous and carbonate depositional environments that relate to petroleum reservoirs are
set forth in Table 1 (Terrigenous depositional environments) and Table 2 (Carbonate
sedimentary environments and their depositional settings) respectively.

Table 1. (Terrigenous depositional environments)

Nonmarine glacial
alluvial fan
braided stream
meandering stream
eolian

Marginal Marine Lacustrine


deltaic
estuarine
barrier island
tidal flat

Marine shelf slope


submarine fan
pelagic

Table 2. (Carbonate sedimentary environments and their depositional settings)

Nonmarine lacustrine
dune
caliche
cave deposits
Coastal beaches
tidal flats (sabkhas in arid regions)
supratidal
intertidal
subtidal, including channel deposits and mud banks
organic swamps

Shelf mud banks


tidal deltas
muddy shelf sands
patch reefs
marine sand bodies
shelf basins

Shelf Margin ecologic reefs


oolitic sand bodies
skeletal sand bodies
patch reefs
beaches
island deposits

Foreslope turbidites
pinnacle reefs
skeletal debris fans and reef talus

Basin euxinic basinal muds


pelagic muds and chalks turbidite sands and debris flows

Depositional Systems
The term depositional system is commonly used to refer to the stratigraphic analog of a major
depositional sedimentary environment (fluvial, deltaic, shelf, etc.). Depositional systems have
been defined as "three-dimensional assemblages of lithofacies genetically linked by active
(modern) or inferred (ancient) processes and environments" (Fisher and McGowen 1967). A
number of contemporaneous or partly contemporaneous depositional systems can be joined
together to form a systems tract (Brown and Fisher, 1977). A typical systems tract could consist,
for example, of fluvial, deltaic, shelf and slope systems, each system genetically related and
gradational to its adjacent system.

Lithofacies
Depositional systems are composed of units of rock termed lithofacies. A lithofacies, often called
"sedimentary facies" or simply "facies," is a three-dimensional mass of sedimentary rock
distinguishable from adjacent units on the basis of its lithology, texture, geometry, sedimentary
structures, and paleontology. A lithofacies is formed in response to the depositional processes
inherent to a specific environment or subenvironment. An active depositional system comprises a
number of depositional subenvironments. For example, a submarine fan may be composed of
slope, feeder channel, channel, interchannel, and other fan lobe subenvironments, each
producing a characteristic lithofacies. Lateral and vertical relationships between individual facies
units may be sharp or gradational.
Depositional Sequences

In a modification of work by Sloss (1963), Mitchum et al. (1977) developed the concept of the
depositional sequence. They define a depositional sequence as "a stratigraphic unit of rock
composed of a relatively conformable succession of genetically related strata and bounded at its
top and base by unconformities or their correlative conformities" (Figure1, Generalized
statigraphic section of a sequence; boundaries defined by surfaces A and B, which pass laterally
from unconformities to correlative conformities). A sequence's boundaries are usually
determined by the physical relationship of the strata themselves (rather than by depositional
processes, rock types, etc.). All of the rocks within a depositional sequence were deposited within
a given period of geologic time. Sequences are composed of a number of bedding surfaces which
represent time lines (chronostratigraphic surfaces). Formations or lithofacies may be parallel to
these chronostratigraphic bedding surfaces or they may cross the bedding surfaces. In the latter
case, lithofacies are termed "time transgressive" (Figure 2 , Generalized cross section across a
deltaic depositional system showing lithofacies transgressing chronostratigraphic time lines).

Bedding units display several physical relationships to sequence boundaries. They may be
discordant, due to an erosional or nondepositional unconformity, or they may be concordant, as in
the case of a nondepositional hiatus (Figure 3 , Various unconformity boundaries of a
depositional sequence). By definition, sequences are bounded by unconformities or their
correlative conformities. Thus at least a portion of the upper and lower boundaries of every
sequence should display some discordance. The discordant upper boundaries (Figure 3 ) may
result either from erosional truncation or toplap. Lower boundary discordance's result from
baselap, which is either on-lap or downlap. Toplap occurs mainly where nondeposition takes
place at the top of inclined strata — for example, where base level equilibrium conditions exist on
a delta plain directly overlying delta-front foreset beds. Onlap occurs where strata terminate updip
against an inclined surface — for example, on the flank of a basin gradually infilling with
sediment. Down-lap occurs where inclined strata terminate against a relatively horizontal surface
— for example, where sediment "starvation" takes place at the distal end of a prograding delta or
submarine fan.

Coastal and shallow marine sedimentation takes place through a combination of lateral
progradation (usually from a preferred direction) and vertical aggradation (Figure 4 , Aggradation
and progradation sedimentary processes). The degree of progradation is controlled by the
amount of sediment influx and the relative position of sea level. Deep marine sedimentation, with
the notable exception of submarine fans, takes place mostly by vertical aggradation of fine-
grained sediment and is not affected by position of sea level. Figure 5 (Transgression, regression
and coastal onlap during relative rise of sea level are of terrigenous influx determines whether
transgression, or stationary shoreline is produced during relative rise of sea level. ) shows the
relative facies and bedding plane relationships for various amounts of sediment influx during a
relative rise of sea level.

Note that, in all three cases of relative sea level rise, stratal surfaces display coastal onlap. With
relatively low influx, transgression takes place and facies climb shoreward through time. In this
case, deeper water facies overlie shallower water facies. When sediment influx is high,
regression takes place and facies climb seaward through time. In this case, shallower water
deposits overlie deeper water facies. When sediment influx and sea level are in balance, facies
may climb vertically.
Figure 6 (Downward shift of coastal onlap indicating relative fall of sea level) shows the
unconformity and downward (seaward) shift in coastal onlap produced by a rapid fall of relative
sea level which takes place between periods of rising sea level.

A relative stillstand of sea level results in coastal deposits not being able to aggrade vertically.
Only horizontal pro-gradation takes place, in this case producing a toplap of stratal surfaces (
Figure 7 , Coastal onlap indicating relative stillstand of sea level).

Environmental Indicators

General Considerations
Figure 1 (Detailed chart showing the five basic parameters and the overall methodology of facies
analysis. Also shown are the types of data normally used to articulate each parameter; these
are listed in vertical order of importance.) presents the five basic parameters and overall
methodology of facies analysis. Also shown, in general vertical order of importance, are the main
sources of data used in defining these facies parameters.

In the early stages of exploration, prior to the clear identification of a particular facies,
explorationists must determine the nature of laterally and vertically associated lithostratigraphic
units. It is thus possible to uncover the genetic relationships between adjacent units that can
provide the interpreter with important clues to depositional settings. As exploration continues —
and more and more data are obtained, say, from boreholes — other parameters, such as
lithology, sedimentary structures, and paleontology, can be more reliably defined.

Traditionally, geometry has been the last aspect of a subsurface facies delineated: data from a
number of boreholes used to be required to adequately define the lateral and vertical extent of a
body of sedimentary rock. The advent of modern seismic techniques, however, such as 3-D and
color-coded seismic displays, allows the geometry of a subsurface facies, in some cases, to be
reliably determined relatively early.
2. LITHOLOGY

Lithology
Lithology, one of the easiest facies parameters to observe, is an important indicator of
depositional environment. This is particularly true of carbonates. Most carbonates are biogenic in
origin and are produced either in situ, or, at most, within a few kilometers of deposition. Thus the
shape, size, and sorting of carbonate sedimentary grains bear a strong relation to the physical
processes that characterize a particular depositional environment. Carbonate sediments are
prone to early diagenesis, including solution, cementation and/or chemical alteration. Certain of
these diagenetic processes are known to develop carbonate porosity. Depositional fabric often
guides and limits this porosity development. Thus, an understanding of depositional environment
can assist the understanding of the diagenetic process leading to reservoir development.

The lithology of terrigenous (clastic) sediment, on the other hand, in addition to being dependent
on depositional environment, is controlled by both its parent rock type and its transportation
history. The depositional fabric of sandstones, however, is easier to detect because they are
generally much less susceptible to diagenesis than carbonates.

Lithologic Parameters of Terrigenous Facies

Grain Size

The grain size of a terrigenous sediment depends primarily on the energy of the depositing
current. Generally, higher-energy currents produce coarser-grained sediments and lower-energy
environments produce finer-grained sediments. Grain size tends to decrease in the direction of
transportation, with the gravel size fraction displaying this feature more so than sand. Sorting
increases along with current energy and the length of time the current is active.
Sedimentary facies commonly display characteristic vertical profiles in which grain size fines
upward, coarsens upward, or remains constant. The determination of such vertical variations in
grain size can be extremely valuable in the diagnosis of depositional environment. Figure 1
(Vertical grain -size profiles of some common sandstone facies) illustrates characteristic vertical
grain-size profiles of some important clastic sedimentary facies. Channel sequences usually
form by lateral accretion and many channel sands display, to some degree, profiles that fine
upward, generally from a scoured base (part a of Figure 1 ). On the other hand, "bar"-type sands
almost always display coarsening-upward profiles, often capped by an abrupt facies change to
fine-grained material (part b of Figure 1 ). Examples include regressive barrier-island sands,
shallow marine bars, delta-front bars, and even eolian dunes. Some facies — for example,
braided-stream sands, tidal sand ridges, and submarine channels — often display "blocky"
profiles whereby grain size remains relatively constant (part c of Figure 1 ). It is important to
remember, however, that vertical variations in grain size can be due to more than one process or
depositional setting; i.e., not all fining-upward sequences are channels, or coarsening-upward
sequences, marine bars. For example, deposition from turbidity currents produces fining-upward
sequences in submarine fans. Also, crevasse subdeltas form coarsening-upward grain-size
profiles similar in appearance to marine bars.

Sands originating in different subenvironments commonly coalesce to form a single vertical grain-
size profile. An example is a deltaic point-bar channel sand that, as it builds outward, incises
down into an underlying delta-front bar sand, forming a deltaic couplet (part d of Figure 1 ). The
surface that divides the two facies is sometimes difficult to determine from well logs, requiring
detailed study of well cuttings and cores. It should be remembered that no single environment
has a unique grain-size profile, and similar profiles may be produced by different environments.
Therefore, profiles should be interpreted with as much supplementary data as possible.
Many attempts have been made to relate the statistical parameters of grain-size distribution to
depositional environments. Brief reviews of the method appear in Reineck and Singh (1980) and
Selley (1985). Three common types of sediment grain-size distribution curves are shown in
Figure 2 (Diagram showing three common types of sediment distribution curves: (1) Frequency
curve; (2) cumulative curve; (3) log-probability curve. Curves (1) and (2) are plotted on an
arithmetic scale; cur e (3) is plotted on a lot-probability scale. Three commonly calculated
parameters Q1 *, Q2 (Md), and Q3 are marked). Grain size is plotted against percent and log-
probability. The more peaked" the frequency curve (1) , and the steeper the cumulative curve (2)
and log-probability curve (3), the better sorted the sample. Since we can compare statistical
analyses of grain size and sorting of ancient sediment with similar analyses of recent sediments
of known depositional setting, it should be possible, by analogy, to determine the ancient
environment of deposition. Yet such attempts have, unfortunately, seldom been successful.

There are several reasons for the failure of this method. A major reason is that grain-size
distribution is a product of hydrodynamic factors. Similar hydrodynamics may be active in more
than one environment, producing similar grain-size distributions in different environments. For
example, fluvial currents in continental meandering stream environments produce sand deposits
with grain-size distributions similar to those in meandering tidal channels.

The history of the sediment may also lead to confusion by not allowing sorting processes to
operate. For example, a sand that has been cleaned and well sorted in a previous depositional
setting may be the only sediment carried into a new environment. Thus only clean, well-sorted
sand can be deposited in the new environment.

Postdepositional mixing of sediment may also obscure results of grain-size analysis. For
example: The clay content of sediment has been shown to be a sensitive indicator of depositional
processes. A clean, well-sorted beach sand may become mixed with clay-sized materials washed
in at a later date from rare flooding in an adjacent river estuary. Yet an ancient sediment may
show no evidence of this subsequent mixing, thus possibly causing a false environmental
diagnosis.

Finally, there is the problem of disaggregating rock samples for use in a sieve analysis.
Postdepositional quartz over-growths and cementation may have completely altered grain sizes
and significantly modified the original texture of a rock.

Chemistry
The significant chemical parameters of terrigenous sediments that may yield environmental clues
are provided mainly by minerals originally precipitated at the time of deposition. These minerals,
however, are extremely susceptible to diagenetic changes and can possibly be more of an
indicator of postdepositional conditions.
The oxidation-reduction potential (Eh) of a depositional environment may be revealed by the type
of iron minerals present; for example, red-colored, ferric oxide cement is formed in an aerated,
oxidizing environment. Drab-colored ferrous oxide, and iron sulfides like pyrite, are formed in a
reducing, oxygen-deficient environment. In general, red-colored sediments indicate continental
deposition in settings above the water table, where acidic, oxygen-rich ground water percolates
through the sediment, precipitating red ferric iron oxides and destroying organic matter. Drab-
colored sediments, green or gray in color from ferrous oxide and preserved organic matter, are
usually indicative of deposition below the ground water table — for example, in meandering flood
plains, coastal environments, and marine settings ( Figure 1 , Sketch showing relationship
between local water table and oxidation. Iron-bearing sediments within the phreatic zone have a
much higher chance of taking on red coloration). The presence of abundant pyrite and
preserved organic matter generally indicates extreme anaerobic conditions in marine or lacustrine
basins where circulation of oxygen-rich water is restricted.

There are some important exceptions to these general statements and using color as a rule of
thumb must therefore be done with caution. As Selley (1985) points out, some deep marine
oozes are red, secondary reddening often takes place under unconformities, and red beds may
become gray-green if they become flushed by strongly reducing connate fluids (commonly
adjacent to petroleum accumulations) .

Mineralogy
Information on depositional environment may be provided by the occurrence of certain common
minerals. For example, glauconite, a greenish-colored complex hydrous alumino-silicate, is an
authigenic (formed in-place) mineral that seems to occur only in marine continental shelf
sediment. It is quite unstable and generally is unable to withstand significant reworking. Its
presence, therefore, has been used with some success as a reliable indicator of deposition in a
marine environment. Its absence, however, does not necessarily indicate a nonmarine
environment.

Another mineral that can be helpful in the diagnosis of depositional environments is detrital mica.
Flakes of mica tend to be winnowed out of high energy environments by turbulence and strong
currents and carried away to be deposited in lower energy environments. Micaceous sands
therefore tend to be absent from well-winnowed environments like barrier islands, shallow shelf
bars, and eolian dunes. Settings in which micaceous sands characteristically occur include outer
delta slopes, outer shelf, and submarine channels and fans.

Particles of carbonaceous material (lignite and coal) are a common minor component of many
sands. Because this material is derived mostly from land plants, it is commonly found in fluvial,
lacustrine, and deltaic sands. Its presence, again, does not necessarily indicate a nonmarine
environment. Carbonaceous material can be found in marine deposits as well. Coal is quite stable
and can survive transport into deep marine systems like submarine fans. Particles of coal are
also frequently encountered in marine limestones, where they probably result from marine algae.
Like micas, carbonaceous debris is easily winnowed away and thus provides, along with mica, a
valuable index of winnowing.

Shell fragments are another common constituent of many sands. Some shelly material can derive
from fresh water organisms. The preservation potential of calcareous particles in many
continental sediments is, however, probably limited, due to leaching by acidic meteoric waters.
The presence of shell material, therefore, most likely indicates a marine depositional setting.
Selley (1985) has shown how the presence or absence of the last four mentioned constituents-
glauconite, mica, carbonaceous material, and shell debris — can be used to define four major
groups of sandstones ( Figure 1 , The presence or absence of glauconite and carbonaceous
detritus divides sands into four main environmental groups: marine winnowed sands; nonmarine
winnowed sands (eolian); mixed sands with both glauconite and carbonaceous matter,
commonly found as turbidites; and poorly winnowed nonmarine sands). High-energy marine
environments like barrier islands, bars, and shoals may contain shelly sands but be without
carbonaceous detritus and mica. Deltaic, fluvial, and lacustrine sands may contain carbonaceous
and/or micaceous material but would not be expected to contain glauconite or shell debris. All
four constituents might be expected to occur in shelf and deep sea marine sands, and be absent
from eolian sands. However, many exceptions exist and, as Selley admits, this technique is "real
cowboy geology," which must never be used in isolation.

Many attempts to use clay minerals as indicators of deposition-al environments have met with
little success, primarily because the chemistry of clays reflects parent rock mineralogy,
weathering, climate, and diagenetic history, as well as depositional conditions. Weaver (1958)
analyzed hundreds of clay samples from ancient rocks and concluded that no single clay mineral
typifies a particular environment. However, his work does show a broad relationship of illite and
montmorillonite to ancient marine rocks and kaolinite to continental rocks.

Krejci-Graf (1972) provides information on the trace element content of sediments in various
depositional environments. Continental sediments that have undergone long periods of subaerial
weathering typically contain the trace elements titanium and thorium. Marine sediments, on the
other hand, are more enriched than continental sediments in the trace elements boron,
chromium, copper, gallium, nickel, and vanadium. Chromium seems to be more abundant in
slightly anaerobic rocks, while vanadium is concentrated in highly anaerobic facies.

Other Lithologic Parameters

Many workers have attempted to use features such as the shape and texture of sandstone as
environmental indicators. Application of these parameters in environmental diagenesis has met
with little success, primarily because the shape and texture of a grain are a function not only of
the most recent depositional process, but of previous history as well. For example, sandstone
grains deposited in a marine or fluvial setting may display a rounded frosted appearance,
probably due to previous transportation in an eolian environment.
3. SEDIMENTARY STRUCTURES, GEOMETRY AND FACIES

Sedimentary Structures
Primary sedimentary structures are formed at the time of deposition, or shortly thereafter, and
prior to consolidation. These primary structures include a great variety of surface markings,
bedforms, bedding plane stratification and penecontemporaneous deformation structures. We
shall concentrate on current-generated stratification features. These are the most common
structures, and generally the most useful in interpreting depositional settings.

Because terrigenous sediments have undergone transportation by currents they are more likely to
possess stratification structures than carbonates, most of which are formed in situ. However,
most carbonate sands (grainstones) have been subjected to significant current activity. Thus, like
their terrigenous counterparts, they may display current-generated stratification features.

Bioturbation structures produced by organic activity also fall under the heading of sedimentary
structures.

A complete review of primary sedimentary structures can be found in Pettijohn and Potter (1964);
Conybeare and Crook (1968); Pettijohn, Potter, and Seiver (1973); Reineck and Singh (1980);
and Collinson and Thompson (1982). For a reference restricted to current-generated stratification
structures in clastic terrigenous sediments, the reader is referred to Harms, Southard, and Walker
(1982). Much of the following discussion is drawn from this source.

Depositional Processes

Before looking closely at flow-produced stratification, let us consider some basic ideas on the
modes of sediment deposition.

Sediment builds up either by vertical accretion, where sediment settles out from suspension, or
lateral accretion, where traction currents supply sediment to a horizontally shifting, sloping
depositional surface. Traction currents, by the way, move sediment by a combination of rolling,
sliding, and saltation (bouncing). In some environments, deposition takes place by a combination
of vertical and lateral accretion. Harms et al. (1982) term this process oblique accretion.

Three basic methods have been recognized by which individual sediment grains come to rest on
a depositional surface: fallout from suspension, traction (with or without fallout), and mass
emplacement.

In fallout from suspension, particles build up beds by vertical accretion. This method
characterizes the lowest current velocities, and generally produces parallel-laminated, fine-
grained deposits in quiet-water lakes and marine basins. It may produce, however, a type of
large-scale cross-stratification in fine sand, termed hummocky cross-stratification.

In tractional deposition, an excess of traction transported grains comes to rest on a depositional


surface that is either planar or rippled (large or small scale). Sediment is supplied from an
upcurrent source by traction accompanied by fallout from suspension. Compared to deposition
from suspension, this type of deposition is usually associated with higher velocity currents and
coarser-grained sediment, although sediment as fine as silt can be deposited in this manner. This
process typically deposits fluvial, coastal, and marine sand bodies that build up by both lateral
and oblique accretion.
In mass emplacement, a high-concentration sediment/fluid mixture propelled by gravity flow
comes to rest to form the sedimentary deposit. Examples include deposition from debris flow and
moving slurries. Grain size varies from clay- to pebble-size and sorting is usually poor. Mass-
emplacement flows include dense sediment-water mixtures called turbidity currents, which move
down steep submarine slopes by gravity. A decrease in gradient causes the flow to lose velocity
and deposit its load in a graded bed, with the coarsest particles settling out first. Mass-
emplacement deposits include debris-flow sands and turbidites in deep water channels and fans.

Bed Configurations

Stratification features are the (partly) preserved remains of flow-produced bed forms. A
classification of these bed forms and a summary of their characteristics as presented by Harms et
al. are listed in Table 1 .

The several basic bed-form configurations are:

• small ripples — downstream-migrating bed forms, produced at relatively low


flow velocities, less than several centimeters in height, and possessing gentle
stoss (upcurrent) and steeper lee (downcurrent) slopes ( Figure 1 , Newly formed
current ripples developing toward equilibrium in fine to medium sand. flow is from
left to right. The top surface of the block is 50 cm x 50 cm).
• two-dimensional large ripples — low (height/spacing ratio), straight crested
features formed at moderate current velocities. They vary significantly in height
from a centimeter to a few meters. Large 2-D ripples are often called megaripples
and the very large forms are called sand waves. They pass gradationally into
three-dimensional forms with increasing current velocity.

• three-dimensional large ripples — similar in profile to small-scale ripples, with


sinuous to highly irregular crest patterns. These forms typically have strong
three-dimensional geometry with uneven crests and troughs. Three-dimensional
ripples are formed in moderate to high current velocities. They range in height
from a few decimeters to a few meters, with a high height-to-spacing ratio
compared to 2-D ripples. In other terminologies they are referred to as
megaripples, dunes, or small sand waves.

• plane beds — flat beds developed at still higher flow velocities in all sizes of
sediment, and also in a narrow range of low velocities for sediment coarser than
about half a millimeter.

• antidunes — waves developed in phase with larger amplitude surface waves at


high current velocities and shallow depths. They are rarely preserved as
sedimentary structures and are not considered here.
The size-velocity diagram in Figure 2 , based on data from numerous flume experiments,
illustrates the fact that bed configuration is strongly dependent on current velocity and, to a
somewhat lesser extent, particle grain size. (Size-velocity diagram for flow depths of 10-22
cm. Open circles, small ripples; plusses, two-dimensional large ripples; solid circles, three-
dimensional large ripples; open triangles, upper plane bed; solid triangles upper plane bed; solid
triangles, antidunes; open inverted triangles, lower plane bed; x marks, no movement.)

Most cross-stratification and cross-lamination features preserved in ancient sedimentary rocks


were produced by the migration of bed forms-current ripples, both large and small. As new
sediment is added and the bed builds upward, ripples migrate across the bedding surface in a
downcurrent direction with a small component of upward climb as well. Figure 3 is a schematic
section through cross-lamination that was produced by such ripple climb. (Sections through
erosional-stoss climbing-ripple cross-lamination, parallel to flow (schematic). Ripple profile
existing at time 1 and 2 shown by heavy lines.) The heavy line shows the profile of a climbing
ripple train. Troughs climb along with the ripples and erosion surfaces are created as the
advancing troughs scour away a portion of the sediment deposited in advance of them. The
dashed segments in the illustration represent parts that are so eroded. The lowest parts of the
foresets are thus preserved as sequences of downcurrent-dipping laminae between parallel
erosion surfaces to form crossbed "sets."

Should no new sediment be added, ripples do not climb. They simply migrate across an area
parallel to the plane of deposition. In this case, the foresets deposited on the lee flank of each
ripple are entirely or mostly eroded away by scour as the following trough goes by.
As shown in Figure 4 , the shapes of foreset laminae deposited by an advancing ripple may
be in the shape of planar laminae, thereby forming a large angle with the underlying trough
surface (part a of Figure 4 ) or they may be sigmoidal laminae, meeting the trough surface
tangentially, at a small angle (part b of Figure 4 ). In cases where the upper portions of the
sigmoidal foresets are scoured away by following troughs, the preserved portion of laminae
assumes a scoop-shaped, concave-upward profile (part c of Figure 4 ).

Flow-Produced Stratification

Stratification produced by steady unidirectional flow includes the following broad categories:

• small-scale ripple cross-lamination;

• large-scale trough cross-stratification;

• large-scale tabular cross-stratification;

• planar lamination.

Stratification produced by oscillatory flow includes:


• small-scale oscillation-ripple cross-lamination;

• hummocky cross-stratification

Small-scale ripple cross-lamination ( Figure 5 , Block diagram showing small-scale ripple cross-
lamination.) occurs as thin sets of mostly concave-upward foreset laminae, dipping in the
direction of local current flow. Each set, in three dimensions, tends to be flat, elongated, and
tongue-shaped. Thickness of sets depends on angle of ripple climb, with high-angle climb
producing sets with the entire ripple profile preserved.

In such cases, sets may reach thicknesses of a few centimeters. Most cross-lamination is formed
by ripples climbing at small angles; sets are thus relatively thin, being only as thick as the
preserved, lower portion of the foreset laminae.
As the name implies, this type of stratification is formed by migrating small-scale, mostly climbing
ripples in sediment ranging from silt to medium-grained sand. Current velocities sufficient for
traction deposition are implied, but velocities are low relative to those generating other types of
current-induced stratification.

Large-scale trough cross-stratification is illustrated by the block diagram in Figure 6 . (Block


diagram showing large-scale trough cross-stratification formed by migration of three-dimensional
large ripples. Flow is from left to right. The length of the sides of the block could range from a few
meters to a few tens of meters.) Set boundaries are approximately parallel, with occasional
wedging out when viewed in a section parallel to flow. From this perspective, these sets could be
mistaken for tabular cross-stratification (described below).

When viewed transverse to flow, however, a trough pattern (originally called "festoon") is clearly
discernible, with sets arranged in a nested pattern, and each set bounded at its base by a
concave-upward erosional surface. Sets are composed of concave-upward foreset laminae
pointing in local downstream direction. Maximum true dips of the foreset laminae vary from 25º to
30º.

Trough cross-strata are formed by the migration of three-dimensional large ripples (dunes and
megaripples) with a small angle of climb in sediment ranging from fine to very coarse sand.
Unidirectional current flow of moderate to high velocity (50 to 100 cm/sec) is implied.
Large-scale tabular cross-stratification ( Figure 7 , Block diagram showing large-scale tabular
cross-stratification formed by migration of two-dimensional large ripples. Flow is form left to right.
The length of the sides of the block could range from ten meters to as much as a few hundred
meters) displays sets with bounding surfaces that are somewhat planar and parallel. The
thickness of these sets may range from a few decimeters to more than 10 meters. Unlike small-
scale cross-lamination and large-scale trough cross-stratification, the foreset laminae are
generally planar as well. Average dip of the laminae is commonly 30º or more, about the
maximum angle of repose. The amount of preserved ripple height may vary considerably,
depending on the angle of climb.

Where similar sets are stacked, one above the other, indicating a train of ripples, climb angles are
probably low and only a small portion of ripple height is preserved. Isolated or solitary sets
developed by large individual sand bodies may have their original ripple height entirely preserved.
Tabular cross-strata are typically produced by migration of two-dimensional large ripples in
unidirectional flows. Current velocities are slightly lower than those producing three-dimensional
large ripples, and they develop in sediment ranging from medium to very coarse sand.

As shown in Figure 7 , smaller ripples, two-dimensional or three-dimensional, are commonly


superimposed on the surface of the large ripples. The strata produced by these small ripples are
not usually preserved, except those reverse-flow ripples formed in the lee of a larger-scale ripple.

Planar lamination ( Figure 8 , Planar lamination produced by aggradation during upper-stage


plane-bed transport) is produced in a wide spectrum of sediment grain sizes by either
unidirectional or oscillatory flow. Currents generating planar lamination occur within two
distinct velocity ranges. An upper-stage regime of strong current velocities (higher than ripple-
forming velocities) produces planar lamination in sediment ranging from silt to medium sand. A
lower-stage current regime within a narrow velocity range of about 40 cm/sec generates planar
lamination, but only in coarse-grained sand.

Environments where upper-stage planar beds may be preserved include stream channels,
beaches, and settings where deposition is by high-speed turbidity currents. Lower-stage planar
stratification is found principally in braided streams.
Cross-lamination from small-scale oscillation ripples, according to Harms et al., are the most
common types of oscillating flow stratification in the sedimentary record and are produced by
migration of climbing ripples where flow is asymmetrical. In this case, laminae usually dip in one
direction only, and stratification geometry is similar to cross-lamination produced by unidirectional
flow small ripples. The closest that stratification comes to having a geometry reflecting pure
oscillatory flow (Figure 9 , Small -scale cross-lamination formed by oscillation ripples) is taken
from a shallow marine sand subjected to wave action.

Hummocky cross-stratification (HCS) is a type of large-scale cross-stratification produced mainly


in coarse silt and fine sand by deposition from suspension on a bedding surface of shifting
hummocks and swales. (Figure 10 , Block diagram showing hummocky cross-stratification.
Current directions unknown.) Hummocks are typically about 10 to 50 cm high and spaced one
to a few meters apart. Although its origin is still somewhat speculative, HCS is generally
interpreted to have been formed on marine shelves, in lakes, and on the lower shoreface by
oscillatory flow from large storm waves at depths below "fair weather" wave base.

Geometry
The overall shape of a lithofacies is controlled by many variables: topography of the original
depositional surface, climate, paleoslope, nature and amount of sediment supply, and intensity
and direction of local depositional currents.
Although facies geometry may be altered by postdepositional tectonic deformation, the use of
appropriate subsurface, isopachous mapping techniques generally allows one to graphically
restore a feature to its original (syndepositional) shape.

Figure 1 (Block diagrams of various facies shapes), illustrates the basic shapes of some major
sedimentary facies. It should be noted, however, that the same geometry may be produced in
more than one environment. Channels may be fluvial, deltaic, tidal, or submarine; fans may be
alluvial, deltaic, or deep marine. A single environment may also produce more than one shape.
Examples include fluvial channels that may vary from straight to sinuous and deltaic river mouth
bars that may assume a variety of shapes, depending on their relative intensities of fluvial versus
marine depositional processes.

Certain generalizations can be made regarding facies geometry. Fluvial channel facies tend to be
elongate-to-sinuous sand bodies, trending perpendicular to the paleoshoreline, and barrier-island
and shelf sand ridge systems develop as elongate shoestring bodies parallel or slightly oblique to
the paleoshoreline. Though a sheet or blanket geometry is generally of little environmental
significance, detailed study may show it to be made up of a series of coalesced braided-stream
channels, barrier-island shoestring deposits, or eolean dune sand bodies.

Geometry, as we have said, has been traditionally the last delineated aspect of a subsurface
facies, since much borehole data were required to define the extent of a body of sedimentary
rock. However, with the advent of modern seismic techniques — such as 3-D seismic and color-
coded seismic displays — the geometry of a subsurface facies can, in some cases, be reliably
determined in the early stages of exploration.
Associated Facies
The identification of certain ancient environments or subenvironments is often difficult. The
environment may not inherently possess enough diagnostic characteristics, or we may not be
able to obtain sufficient data for positive identification. In such cases, the analysis and
identification of associated environments become important tools in the interpretation of the facies
at hand.

With the rise and fall of sea level, shorelines transgress landward or regress in a seaward
direction. Environments migrate laterally over one another, depositing a conformable succession
of genetically related sediments with both systematic and predictable gradational facies
transitions. These gradations take place in both vertical and lateral directions ( Figure 1 , Facies
relationships resulting from marine transgression and regression). Similar gradations develop
between the subfacies of a depositional system, as subenvironments shift specially with time.
Lateral migration of subenvironments, though, may also take place in response to purely local
factors, such as delta progradation or stream migration. This critical relationship between
environments and facies provided the basis of Walther's (1894) law of succession that states "...
only those facies and facies areas can be superimposed that can be observed beside each other
at the present time." This principle forms the foundation of an important method in facies
identification — analysis of the environments of the associated facies above and below, as well
as laterally, allows us to reconstruct the most logical sequence of depositional events.
4. PALAEONTOLOGY

Microfossils
Microfossils are by far the most abundant of all fossils in the paleontological record. Microfossils,
moreover, generally survive the action of the drill bit, and quantities of them are retrieved from
well cuttings, whereas macrofossils become fragmented and are rendered generally
unidentifiable by drilling operations. Microfossils are most often found in the cuttings of fine-
grained sediments like shales and limestones, but are relatively rare in sandstone. Because
various types of microorganisms flourished in different depositional settings, their fossil remains
provide extremely valuable indicators of sedimentary environments.

The other major application of microfossils is in biostratigraphy, as chronometers of geologic time.


Faunal evolution throughout the Phanerozoic (Cambrian to present) has allowed the ages of rock
samples to be determined and strata to be correlated by the recognition of the distinct species
and assemblages of microfossils they contain.

The two most important varieties of microfossils employed in environmental analysis are
foraminifera and ostracods. The main reason for this importance is that some suborders and
genera of foraminifera and ostracods are benthic (bottom dwellers) and live only in specific
environments according to depth and salinity. Thus, when found as an assemblage of fossils in
the same place they lived, termed a biocoenose, they can provide a direct indicator of
depositional setting.

Foramenifera
Foraminifera ("forams") are the most widely used microfossils. These single-celled organisms
have existed since the Ordovician, with 1400 genera and 30,000 species (4,500 still in existence)
having been classified.
The skeleton, or test, of foraminifera averages 0.5 mm in size and is distinguished by chambers
interconnected by an opening, or foramina ( Figure 1 , Chamber construction in Dicorbis).
Although they have a simple structure, foraminifera tests come in a wide variety of shapes
(Figure 2 , Variety of shapes of foraminifera). Foraminifera, both planktonic (floating) and
benthic (bottom-dwelling) forms, inhabit a wide variety of aquatic environments, from very shallow
water to depths of 5000 in, and in waters ranging from brackish to hypersaline. They are not
found, however, in fresh-water lakes.

It has been determined that the habitat of recent foraminifera depends largely on the composition
of their tests. Three basic groups of test composition are recognized:

• agglutinated — tests formed from material borrowed from the habitat: grains of quartz
(arenaceous tests), flakes of mica, clay material, and various skeletal debris. The material
Is bonded together by a secreted cement, either chitinoid (celluloselike) or calcitic. The
relative proportions of cement and agglutinate vary with species.

• porcelaneous — opaque, calcareous tests that appear white and brilliant in


reflected light.

• hyaline — calcareous tests characterized by their glasslike transparency.

Salinity. Most benthic foraminifera are marine. However, certain groups having porcelaneous tests live
equally well in hypersaline lagoonal environments. Other types, such as certain agglutinates and hyalines
prefer brackish lagoons and estuaries. Some foraminifera may be found in all environments.
Water Depth. Figure 3 (Depth distribution of recent benthic froaminifera) shows the depth
distribution of recent benthic foraminifera. Generally, benthic foraminifera with porcelaneous
tests live in shallow waters, whereas hyaline tests occur everywhere but in deepest waters.
Agglutinated types similarly occur everywhere, with noncalcareous benthic types surviving at
great depths (up to 5000 in).

A fourth group of foraminifera, characterized by microgranular tests, flourished in the upper


Paleozoic but are now extinct. Figure 4 (Stratigraphic range of some foraminiferan groups) 1
illustrates the stratigraphic range of the families within the four main groups.

The standard reference on foraminifera is Loeblich and Tappan (1946). More recent works
include Haynes (1981) and Buzas and Sen Gupta (1982).
Ostracodes

Ostracods are the most advanced forms of microfossils used. They are crustaceans and can be
regarded as shrimps within a bean-shaped calcitic shell ( Figure 1 , Lateral view of a podocopid
ostracod, without the left valve: A1 - 7 = appendages; a = anus; c = carapace; dg = digestive
system; e = eye; f = furca; go = genital organs; m = mouth (x100)). When the ostracod dies,
the body disappears, leaving the shell (carapace) to be fossilized ( Figure 2 , Photo micrographs
of ostracod shells). Ostracods have existed since the Cambrian; their fossils abound in sediments
of all kinds but are particularly abundant in clays and marls. Shell size ranges from 0.15 to 20
mm for living species and up to 80 mm for specimens found in Paleozoic rocks. Ostracods inhabit
every aquatic environment, both marine and fresh water. Many varieties are benthic and prefer
quiet, still waters where finegrained sediments are rich in organic material. Basically three broad
assemblages are recognized based on sensitivity to salinity:
• fresh water — generally with thin, smooth shells.

• marine — robust shells, often ornamented.

• brackish water — various shell markings, depending on salinity.

Assemblages of marine benthic varieties can also be used to give marine water depth; shelf (less than 200
in), bathyl (200-500 m), and abyssal (greater than 500 in).

For additional information on ostracods, the reader is referred to Van Morkhoven (1962-3) and
Bate, Robinson, and Sheppard (1982) .

Other Common Microfossils


As summarized by Bignot (1985), other common microfossils, some of which have value as
environmental indicators, include calcareous nanofossils, radiolarians, diatoms, conodonts,
spores and pollens, and dinoflagellates.

Calcareous nanofossils — minute (1 to 35 micron) rosette-, star-, or button-shaped plates called


coccoliths — are the most common variety ( Figure 1 , Discoasterid calcareous nannofossils and
other presumed coccoliths). Because of their small size, they are best studied with the
electron microscope. Fossil coccoliths, common since the Jurassic, are found mostly in marine
sediments deposited far from coasts. No coccoliths have been found in lacustrine sediments.
Radiolarians — marine planktonic microorganisms, whose preserved siliceous skeletons consist
of needlelike spicules — are found in rocks ranging from the Ordovician to Recent; the presence
of radiolarian microfossils indicates marine deposition in waters of normal salinity ( Figure 2 ,
Different types of siliceous radiolarians).

Diatoms are single-celled algae whose siliceous microfossils date from the Late Cretaceous (
Figure 3 , Different types of diatoms). They are found in all types of aquatic environment
regardless of salinity; many varieties are benthic, and most species are subject to strict ecological
controls. Fossil diatoms should thus be expected to Provide information on depositional
environments.
Conodonts are small structures, averaging 1 mm in length and consisting of calcium phosphate,
that appear to be the teeth or hooks of larger organisms

( Figure 4 Diagram of a composite conodont). They occur only in Paleozoic and Triassic marine
sedimentary rocks, both in near-shore and glauconitic sands and in deeper, fine-grained facies.

Spores and pollens originate from higher land plants, and generally indicate continental
paleoenvironments. However, because they are subject to wide dispersal by both wind and water
currents, they may also be found in a wide variety of marine deposits. Common since the
Devonian, spores and pollens provide excellent biostratigraphic markers.
Dinoflagellates are single-celled algae whose fibrous cellular walls are preserved as fossils,
termed cysts ( Figure 5 , Different types of dinoflagellates). Dinoflagellate cysts range in size
from 60 to 150 microns, and in age from Permian to Recent. Dinoflagellates occur as plankton in
the surface waters of all environments ranging from marine to lagoonal to lacustrine. Because of
their wide dispersal, they are not reliable indicators of depositional environment.

By way of summary, Figure 6 (Simplified classification of aquatic environments and salinity


ranges for some groups of living microorganisms) illustrates the salinity ranges for some common
groups of living organisms that provide fossils in the geologic record.

Trace Fossils

Trace fossils constitute another valuable type of biological information that can be used in
environmental analysis. Subsurface information on trace fossils, however, may not always be
available. Unlike microfossil data easily obtainable from well cuttings, subsurface information on
trace fossils can only be derived from expensive conventional cores. Trace fossils, produced by
bioturbation, are simply fossilized tracks, trails, burrows, etc. made by animals within sediment or
on the sediment surface. They provide a reliable record of benthonic organic communities
because, unlike other fossils, they always occur in situ and cannot be reconcentrated by
reworking. These structures also include plant trace fossils, like root molds and casts.
Seilacher (1967) devised a series of ichnofacies based on the observation that certain types of
animal trace fossils characterize particular environments (Figure 1 , characteristic ichnofacies for
various environments). Each ichnofacies consists of a suite of trace fossils which can be used
as an indicator of bathymetry.

Skolithos ichnofacies consist of vertical burrows made in sandy or firm mud bottoms of the littoral
(intertidal) zone by suspension feeders, i.e., by organisms that feed on food particles suspended
in the agitated zone of shallow water.

Cruziana and zoophycos ichnofacies consist of more horizontal and increasingly patterned
burrows made in soft muds of the neritic zone by sediment feeders.

Nereites ichnofacies consist of crawling traces made in the soft muds of the bathyl zone by
sediment-feeding organisms.

Although the animals responsible for the trace fossils have undergone evolution, these
ichnofacies have remained remarkably constant through Phanerozoic time.
5. SOURCES OF INFROMATION

General Considerations
The analysis of any sedimentary environment depends on two basic types of information. One is
derived from direct observation of the rocks themselves — from outcrops, cores, and well
cuttings. The second type is indirect, being generated by instrumentation: wireline logs (including
dipmeter) and seismic surveys.

The most valuable direct type of information is, of course, derived from outcrops, particularly
direct data on large-scale sedimentary structures that are, otherwise, generally unobtainable. The
petroleum geologist is, however, primarily concerned with the subsurface and therefore key
stratigraphic horizons may not be exposed at the surface. Even if exposed, multiple facies
changes may result in the outcrop information being of little value for the problem at hand.

Cores and Cuttings


Most rotary drilling bits provide cuttings large enough and fresh enough in appearance to allow
lithological evaluation. It is possible, therefore, while drilling is taking place, to prepare an
accurate lithological section of a well called a sample log. There are several methods for
preparing a geological sample log. The most important of these for facies analysis is a log that
provides a complete written description of each sample ( Figure 1 , Example of a well sample
log). Important formation boundaries are noted, and a symbolic lithological column useful for
quick reference is often included.

At the wellsite the log is compiled by a geologist or mudlogger, and, naturally, the resulting log
is only as reliable as the experience and qualifications of the person preparing it. Wellsite sample
logs are often Part of a mudlog, which contains abundant information on the condition of the
drilling mud, hydrocarbon shows, drilling rate, etc. Often sample logs prepared at the wellsite
contain insufficient data for purposes of environmental analysis, geologic dating, stratigraphic
correlation, paleontology and palynology, and so on. In such cases, experienced geologists
should reexamine samples and compile detailed logs, complete with facies interpretations, where
possible.
Cores provide the other source of direct data from boreholes. There are two methods of obtaining
core samples: (1) from a drilled interval by bottomhole conventional coring, done during drilling
and (2) from sidewall coring, accomplished after drilling. The conventional bottomhole cores are
taken by a special device called a core barrel, which replaces the normal drill bit at the bottom of
the drillpipe assembly ( Figure 2 , Core barrel operation: (a) Running into the hole, (b) circulating
on bottom, (c) Coring). Cores up to 5 in. (12.3 cm) in diameter and up to 90 ft (28 in) long are
obtained. They provide excellent lithological and paleontological samples, often continuous and
undisturbed.

Sedimentary structures are frequently displayed and valuable information on the nature and
direction of paleocurrents can thus be obtained, particularly if the position of the core with respect
to magnetic north is recorded. Coring equipment is expensive, and coring proceeds at slow rates,
thus adding to rig-time costs.

A geologist, therefore, must provide strong justification when recommending a conventional


coring operation to management. Sidewall coring, carried out after drilling, is completed by means
of a core gun on a wireline logging cable. At the appropriate sample depth, a 1-in (2.5-cm) hollow
bullet is fired horizontally into the borehole wall and a core up to 2 in (5 cm) long recovered
(Figure 3 , Sample-taking operation of the sidewall gun). Sidewall cores are small but
relatively cheap, and can be obtained selectively over an extensive depth range. Although often
damaged by impact, they can provide useful data on lithology and paleontology for environmental
analysis.

The description of cores and cuttings should follow some kind of standardized order. A scheme
published by the American Association of Petroleum Geologists in the "Sample Examination
Manual" (Swanson 1981) is recommended:
• rock type — should be underlined and followed by proper compositional or
textural classification terms, e.g., "limestone: fossiliferous grainstone;"

• color;

• texture — including grain size, shape, sphericity, roundness, and


sorting;cement and/or matrix materials — cement is deposited chemically around
the grains, and includes silica and calcite; matrix consists of small individual
grains that fill interstices between larger grains, and includes clay and silt;

• fossils — includes microfossils, small macrofossils, and fragments of fossils;

• accessory constituent — minerals, rock fragments, and plant debris that


constitutes only a minor percentage of the bulk of the rock, e.g., glauconite, mica,
carbonized plant remains, and sand-size rock fragments;

• sedimentary structures — most not discernible in cuttings or sidewall cores but


can be found in conventional cores;

• porosity and permeability;

• hydrocarbon shows.

Conventional abbreviations are used in sample descriptions. These abbreviations, along with a
description of sampling techniques, charts, and conventional legends and classifications, can be
found in the AAPG manual by Swanson (1981) mentioned earlier.

Wireline Logs
Wireline logs provide two basic functions. They provide both the data for evaluating the
hydrocarbon-bearing Properties of a zone (formation evaluation) and the control for subsurface
mapping. In formation evaluation, logs are used to define physical rock characteristics, such as
lithology, porosity and permeability; to distinguish between oil, gas, and water in the reservoir;
and to estimate reserves. In subsurface mapping, logs are used to correlate zones, to construct
cross sections, and to provide control for structure and isopach maps.

There are, in addition to the above functions, two very important uses of well logs in facies
analysis: as direct indicators of vertical grain-size profiles by spontaneous potential (SP) and
gamma ray curves, and in interpretation of sedimentary structures by the dipmeter log. Used
together, they can be a powerful tool in environmental diagnosis.

Interpretation of Grain-Size Profiles from Well Logs

Certain types of sedimentary facies display characteristic grain-size distribution profiles. These
profiles may be revealed on spontaneous potential (SP) and gamma ray logs. The SP log records
the voltage differences between an electrode move along the wellbore and the potential of a fixed
electrode at the surface. This potential response to electrochemical factors within the borehole is
brought about by differences in salinity between the mud filtrate and formation water within
permeable beds. These factors are essentially related to the permeability of the bed.
A major factor in the reduction of permeability in a formation is the presence of shale. The SP log
response is thus a measure of shale content. Because the amount of shale matrix in most
sandstones tends to increase with decreasing grain size, the SP log can be used as an indicator
of vertical grain-size variations. The SP curve, measured in millivolts and recorded on the left-
hand side of the log display, varies between two extremes — a shale baseline and a line
corresponding essentially to clean sand ( Figure 1 , Example of SP log in a sand-shale series).
As grain size increases, the curve tends to deflect to the left, toward the clean sand line; as it
decreases, the curve deflects to the right, toward the shale baseline. SP deflection may be
reduced in permeable beds if any are thin or contain hydrocarbons.
The above relationships hold true only for thick beds with for-mat ion water resistivities less than
the drilling mud. Formation waters more resistive than the drilling mud — for example, fresh water
in the formation — will cause a reversal of SP deflection to the right side of the shale baseline (
Figure 2 , SP curve in sand for various thicknesses and formation water salinities). Because a
conductive fluid is required in the borehole, the SP will not operate when nonconductive drilling
muds, such as oil-base muds, are used.

Use of the SP log as a vertical grain-size profile is valid only for sediments with primary
intergranular porosity. Thus, it is generally not a reliable indicator of vertical grain-size distribution
in cemented sandstones or most carbonates.

The second wireline log used to obtain vertical grain-size profiles is the gamma ray. Gamma ray
logs measure the natural radioactivity of formations. Shale-free sandstones and carbonates
usually have low concentrations of radioactive materials, whereas shale has relatively high
concentrations of the radioactive elements uranium, potassium, and thorium. The gamma ray log
is thus used to estimate the amount of shale in a formation. The gamma ray curve, like the SP
curve, is recorded on the left-hand track of the log display and records high concentrations of
radioactivity by deflection of the curve to the right ( Figure 3 , Example of a gamma ray log).

As mentioned earlier, the amount of shale in a formation tends to increase with decreasing
grain size. Therefore, as in the case of the SP curve, deflections of the gamma curve to the right
normally indicate decreasing grain size.

The gamma ray log, like the SP log, has its limitations. Clean, shale-free sandstone may produce
a high gamma-ray reading if it contains potassium feldspars, micas, glauconite, or uranium salts.
The high readings produced in such cases can make a clean sand appear fine and shaly.
Conversely, kaolin-and chlorite-rich shales, because of their low potassium content, may produce
lower than normal gamma readings.
As pointed out, no single environment displays a completely unique grain-size profile. Thus
environmental interpretation of SP/gamma ray curves should take into account as much
supplemental data as possible. Selley (1985) presented environmental interpretations for four
basic SP/gamma log profiles that depend on the presence or absence of glauconite, shell debris,
carbonaceous detritus and mica. ( Figure 4 , Four characteristic gamma log motifs. From left
to right: thinly interbedded sand and shale; and upward-coarsening profile with an abrupt upper
sand-shale contact; a uniform sand with abrupt upper and lower contacts: and, furthest right, an
upward-fining sand -shale sequence with an abrupt base. None of these motifs is environmentally
diagnostic on its own. Coupled with data on their glauconite and carbonaceous detritus content,
however, they define the origin of many sand bodies.)

Use of the Dipmeter in Facies Analysis

The standard dipmeter tool is a wireline logging device consisting of micro-resistivity electrodes
mounted on four pads equally spaced at 90º from one another. The tool is gradually raised
through the borehole and the readings from each of the four pad electrodes are recorded as
resistivity curves. A recording is also made of the tool's position relative to magnetic north.

A resistivity anomaly is usually produced by a bedding plane intersecting the borehole, the
character of the anomaly being roughly similar on each of the four resistivity curves. A computer
correlates the four curves and calculates the vertical displacement of one curve to another (
Figure 5 , Mode of operation of the dipmeter log showing how dip directions are calculated from
the four mutually opposed resistivity curves). The dip angle and azimuth of the bed are then
computed and presented on one of several displays.
The most common of these displays is the arrow "tadpole" plot ( Figure 6 ). On a typical plot,
dip is read by the position of the tadpole base on the dip scale and the azimuth is read by the
direction in which the tadpole tail points.

In addition to its obvious importance in diagnosing structural characteristics, such as folds,


faults and unconformities, the dipmeter can be extremely valuable in facies analysis, particularly
as an indicator of sedimentary structures. It has been found on tadpole plots that dips arrange
themselves into characteristic patterns. When reflecting sedimentary structure these patterns,
termed depositional patterns, consist of three basic types: slope patterns, current patterns, and
low-energy structural patterns. Combined with SP/gamma ray profiles these patterns become
extremely valuable indicators of depositional environments.

Slope patterns are characterized by upward-decreasing dips (red dip pattern) generally having a
common direction. When generated within sandstone they usually represent lateral accretion
surfaces of a channel sandstone. ( Figure 7, Idealized dip log pattern showing progressively lower
slope amount (red motif) characteristic of filled-in channels. Tadpoles shown correspond to
dips of major accretion surfaces - in this case, those of the point bar. Note vertical exaggeration
of cross section.) Such dips point in the direction of the stream channel and perpendicular to
stream flow.
Slope patterns may also be developed in fine-grained sediments where they represent drape or
differential compaction over more rigid underlying features, such as sand bars or reefs ( Figure 8 ,
Red pattern on dipmeter resulting from differential compaction of shale over underlying rigid
feature). These dips point in a direction away from the crestal high of the underlying feature
and are really more structural than depositional in origin.

Current patterns are upward-increasing dips of common direction (blue patterns) generated by
the concave-upward foresets of current-induced cross-stratification. They naturally point in a
downcurrent direction. Because of the limited thickness of many individual cross-strata sets,
recognition by the dipmeter often requires use of computer programs that calculate dip in very
small vertical intervals. ( Figure 9 , Dip patterns related to current bedding produced by westward
current flow. Examples C, D and E illustrate the results of using a 2-ft correlation interval in beds
of varying thickness.)
Upward-increasing blue patterns are also produced by prograding deltas, barrier-island
sequences, and submarine fans. In these cases, dip generally increases upward along with
increasing grain size, and a single pattern may extend over a large vertical interval.

Low-energy structural patterns are generally low-angle, parallel dip (green patterns), typically
occurring in shale. In addition to their presence in vertically extensive shale sequences, they
occur in shale units interbedded within sand bodies ( Figure 10 , Common dip patterns and
coloring code). Most shale is assumed to have been deposited on essentially flat, horizontal
depositional surfaces. Therefore, any green pattern dips over two degrees or so are likely to
represent postdepositional structural tilting.

Use of Porosity Logs as Indicators of Lithology

Much useful information on lithology can be gathered by using combinations of conventional


porosity tool measurements. The most useful combinations are:

• crossplots such as bulk density versus neutron porosity, bulk density versus
sonic travel time, and sonic travel time versus neutron porosity.

• M-N and MID plots, whereby three log readings (neutron density and sonic) are
reduced to two-dimensional crossplots.
It is possible to scale porosity logs so that two curves, when overlain and compared with a
gamma ray curve, immediately give a visual indication of rock type. Figure 11 (Example of
generalized lithology logging with combination gamma ray neutron (CNL)-density (FDC) log)
shows how a combination gamma-ray, neutron-density log can be used as a tool for determining
lithology. Figure 12 (Example of a combination gamma ray (GR) neutron (N)-density (d) log
showing corresponding lithologies from the Ordovician Red River formation, Richland County,
Montana ) is a combination gamma-ray, neutron-density log showing corresponding lithologies
within a carbonate sequence in the Williston Basin of Montana.
Figure 13 illustrates a gamma ray spectral log. Unlike the gamma ray log, which measures
total radioactivity (left tracks), the spectral log reads the relative concentrations of radioactive
potassium, thorium, and uranium (right tracks). The thorium-uranium ratio measured by this log
has been found to be a valuable indicator of depositional environment (Fertl 1979).

A thorium-uranium ratio greater than 7 is thought to indicate a continental, oxidizing environment


and a ratio of less than 7 to imply marine deposits, most likely gray and green shales. For
thorium-uranium ratios less than 2, the presence of black, probably organic, shales deposited in
anoxic marine environments is suggested. For example, at point "A" on the log in Figure 13 , the
thorium curve reads about 14 ppm and the uranium curve about 8 ppm, yielding a thorium-
uranium ratio of 1.75. Thus, a black marine shale is indicated.

Gamma Ray Spectral Log

The gamma ray spectral log may also be used for lithological identification, particularly for clay-
typing. The crossplot chart in Figure 14 (Thorium/ potassium crossplot for minerals identification )
maps a number of radioactive minerals according to their thorium and potassium concentrations.
Again, looking at point "A" on the log in Figure 13 , we see that the thorium curve reads about
14 ppm and the potassium curve reads 2.5%. Applying these readings to the crossplot in
Figure 14 , a clay of mixed-layer composition is indicated.
Seismic
A very valuable source of information in environmental analysis is provided by seismic data. The
importance of this type of data has increased with the development over the past 20 years of a
new branch of seismic interpretation known as seismic stratigraphy. Papers describing the
principles of this method were assembled in the landmark volume AAPG Memoir 26, edited by
Payton (1977). A concise and readable treatment of the concepts and procedures presented in
this volume can be found in Sheriff (1980).

Two fundamental procedures in seismic stratigraphy that bear on the analysis of sedimentary
facies are:

• seismic sequence analysis — the delineation of depositional sequences by


analysis of seismic data.

• seismic facies analysis — the examination of reflection character to aid in


defining depositional environment.

Seismic Sequence Analysis

Seismic sequences are the seismic expression of depositional sequences. They are a package of
seismic reflections from sediments deposited within a specific time-stratigraphic unit. The
isochronous stratal surfaces within sequences and the unconformities and correlative
conformities defining sequence boundaries generate the reflections. Sequence boundaries are
characterized by both discordant and concordant reflections, with the discordant reflections being
either erosional (truncated) or depositional (lapouts) (Figure 1 , Relation of reflections within a
sequence unit to the unit boundaries. (a) Relations at top of sequence unit. (b) Relations at
base of sequence unit. (c) Reflection terminations within an idealized unit.).
A crucial point to remember in seismic sequence analysis is that seismic reflections generally
indicate time surfaces, not necessarily surfaces of lithofacies. An example of this relationship, as
presented by Vail et al. (1977), is illustrated in Figure 2 (Correlation between two wells) and
Figure 3 (Correlation between two wells shown in Figure 2 after more information is added. A
number of other wells intervene between wells 5 and 1). A line connecting the top of a basal
transgressive sand encountered in two wells ( Figure 2 ) represents a lithofacies surface.
However, intervening control (not shown) indicates that the basal sand in the left well is older than
that in the right well, and that time (stratal) lines cross the facies line ( Figure 3 ).
Seismic Facies Analysis

A seismic facies is a three-dimensional unit of reflections whose characteristics differ in some


respects from adjacent reflections. Seismic facies distinctions are often used, in addition to
unconformities, to define seismic sequence units. Thus the terms "seismic sequence" and
"seismic facies unit" are sometimes used interchangeably. However, we should try to distinguish
between them. As previously mentioned, a seismic sequence is a package of reflections resulting
from sediments deposited within a time-stratigraphic depositional unit, whereas a seismic facies
describes the character and configuration of seismic reflections resulting from the depositional
environment. A seismic facies is thus the sonic response to a lithofacies.
Factors that distinguish one seismic facies from another include reflection configuration,
continuity, and amplitude, as well as the geometry of the reflection package (sequence unit). One
type of internal configuration particularly useful in environmental reconstruction involves
progradation (outbuilding) and can be divided into two classes, oblique and sigmoid
( Figure 4 , Sigmoid and oblique progradational types). Oblique progradational units are
characterized by toplap or angularity at the upper boundary of the sequence. This type of
configuration implies high-energy depositional conditions, for example, at the front of a prograding
delta. A fairly clean sand can thus be anticipated at the top of oblique patterns. Sigmoid
progradational units are distinguished by flattened S-shaped reflections indicative of upbuilding as
well as outbuilding of the sedimentary prism. Here, deposition usually takes place in quiet, deep
water and sediments are usually fine grained. A prograding carbonate shelf typically generates
this type of reflection pattern.

Reflection amplitude and reflection continuity are also particularly useful in environmental
diagnosis. High amplitude reflections with good lateral continuity characterize sediments
deposited in quiet water conditions, e.g., interbedded marine limestones and shale. ( Figure 5 ,
The strength and excellent continuity of reflection B suggest the contact between a shale and a
limestone, both deposited under low-energy conditions.)
High amplitude reflections with poor continuity imply continental deposition, e.g., inter-bedded
channel sands and shales ( Figure 6 , In the absence of tectonic complications, poor continuity
suggests nonmarine sediments). Shales and siltstones typical of deeper water settings tend to
be thin bedded and produce relatively closely spaced reflections with good continuity (part a of
Figure 7 , Seismic examples of fine-grained sediments showing contrasting reflection
characteristics. ). Reflection amplitude tends to be moderate to poor depending on lithology
and bed spacing. However, acoustic impedance contrasts can be so low in fine-grained
deepwater sediments that they appear almost reflection free (part b of Figure 7 ) Hummocky
reflections can be produced in high energy, shallow water conditions, e.g., by shallow marine
sand bodies ( Figure 8 , Hummocky form often characteristic of deposition in active shallow
water).

The three-dimensional shape of sequence units often provides an indication of depositional


setting, as shown by the shape of the basin-slope and basin-fill units in Figure 9 . (Basin slope
and basin-fill seismic facies types )
Facies analysis can be further aided by the use of special seismic techniques, such as 3-D and
color displays. There can be no more obvious indication of a depositional environment than that
presented by the meandering stream pattern on the 3-D horizontal time slice in Figure 10
(Horizontal time slice showing the meandering channel geometry and oxbow lakes of an alluvial
flood plain ).

Color displays of reflection amplitude, acoustic impedance, frequency, and velocity can often
reveal facies changes and environmental relationships that are not evident on conventional
seismic sections.

For full descriptions of the many reflection characteristics and their environmental significance,
the reader is referred to the above-mentioned works (Payton 1977 and Sheriff 1980).
6. SANDSTONE RESERVOIR FACIES

MAJOR SANDSTONE RESERVOIR FACIES

The key parameters exhibited by important sandstone reservoir facies and some of the
associated facies are listed in the following pages. Also, relevant diagnostic evidence — in the
form of cores, logs, and seismic — is summarized for cases where it is deemed particularly
relevant.

6a. Alluvial Fans


Generally, the style of deposition on alluvial fans prevents them from acting as good reservoirs.
To date, there are only a few, clear-cut examples of fields producing from terrestrial fan facies.
They are often, however, extremely important to recognize and delineate in the subsurface
because of their indication of both tectonic setting and source area composition. Given this, and
the fact that fans commonly grade into fan delta and alluvial plain environments (whose
sediments have far greater potential to be good reservoirs), alluvial fans can serve as associated
facies of crucial significance.

Summary of Facies Characteristics

Lithology

• fanglomerate (some very large fragments)

• channel sand, conglomerate

• thin shale layers

• rapid vertical and lateral changes

• commonly red beds

Sedimentary Structures
• crude to unbedded (fanglomerate)

• imbricated and oriented pebbles

• crossbedding in channels (various scales)

• crude horizontal stratification

• current lineations

Paleontology
• rare vertebrate bones, plant debris

• more common spores, pollen, often oxidized


Geometry
• fan-shaped in plan view

• wedge-shaped in radial profile

• convex-upward in transverse profile

Associated Facies
• fault-generated mountain fronts

• mountain stream valley

• alluvial plains (braided river)

• playas and eolian facies

Diagnostic Evidence

Cores

The general coarseness, poor sorting, clast angularity, and immaturity of alluvial fan sediments
are the conspicuous features that dominate most core samples of the upper fan. Finer-grained,
cross-stratified or flat-bedded channel sandstones can also be prevalent, particularly from lower-
fan sediments. Figure 1 (Devonian alluvial fan successions in the Hornelen Basin, Norway )
shows the internal details of three alluvial fan coarsening-upward sequences from the Devonian
of Norway.
Logs

Figure 2 (Idealized gamma ray and dipmeter logs for an alluvial fan sequence, showing both
fanglomerate and channel development. Note three major patterns: lowest green dips
represent shale breaks and correspond to spikes on gamma ray curve; random "bag o' nails" dips
in fanglomerate; and dip clusters that show an upward-increasing blue pattern in channel sands )
shows an idealized gamma-ray/dipmeter profile through a fan. The log is characterized by a
monotonous gamma ray curve generated by fanglomerate and coarse braided channel sand.
Several shale layers are indicated simultaneously by spikes of high radioactivity on the gamma
ray log and low dips (green motif) on the dipmeter log. These shales separate three channels,
whose tadpole patterns show a clustering of high-angle dips caused by crossbedding.

Seismic
Vertical seismic sections through alluvial fan complexes typically show discontinuous internal
reflectors. ( Figure 3 , Seismic section and interpretation through probable alluvial fan developed
over structurally deformed basement. Note relatively poor internal seismic character of the
deposit.) This should be expected, given the great lateral and vertical variation in lithology.

6b. Braided Streams

Many present-day alluvial fans pass laterally into the alluvial plain of a braided river, that is, one
characterized by an interlacing veinlike network of low-sinuosity channels with constantly shifting
midchannel bars ( Figure 1 , Block diagram model of a braided stream system in a semi-arid
environment). Streams and rivers tend to braid when three main factors conspire: (1) high
(though possibly seasonal) discharge, (2) relatively steep slopes and (3) large amounts of coarse
bedded sediment.

Summary of Facies Characteristics

Lithology

• highly variable

• up to 90% coarse, pebbly sandstone

Sedimentary Structures
• soft sediment deformation

• ripples

• planar bedding

• tabular and trough crossbedding


• crude bedding, oriented pebbles

Paleontology
• some plant and animal debris, highly oxidized

• rootlet horizons

• burrows

Geometry
• sheetlike, may cover thousands of square miles and be hundreds of feet thick

Associated Facies
• proximal: alluvial fan

• distal: meandering stream alluvium, sabkha, eolian dunes, playa (desert lake),
possible transition to marine delta

Diagnostic Evidence

Cores

Core samples taken from a braided alluvial section can reveal either a homogeneous section of
coarse, crossbedded and gravelly sandstones or a diverse range of grain sizes and sedimentary
structures. Again, well-preserved individual sequences begin with a sharp erosional base that
marks the channel floor, possibly overlain by an upwardly fining progression of grain sizes and
sedimentary structures. ( Figure 2 Idealized "outcrop" showing succession of grain sizes and
sedimentary structures in a single channel sequence of braided alluvial system.)

Logs

Figure 3 (Log of braided alluvial sequence showing characteristic monotonous log response.
Note that the gamma log is neither as clean as that for eolian deposits, not as shaly as
meander channel flood plane alluvial sequences. Azimuth frequency plots reflect linear trend of
this river type ) displays the idealized log response of a braided stream deposit. Some crude
fining-upward portions of the curve can be discerned, but grain-size variation is most often too
small to produce a convincing bell-shaped channel profile, and blocky profiles usually result.

In terms of its dipmeter signature, this facies mainly shows the multiple stacking of channels.
Within each channel, azimuths and dip amounts are clustered into separable groupings. Channel
switching is characteristic but azimuth changes usually remain within a 90º arc. The probable
long dimension of the sand body as a whole can often be found by bisecting the arc when it is
plotted on an azimuth frequency diagram.

Seismic
Due to relative lithologic homogeneity, braided stream deposits do not often show internal
reflections. Shales are too thin and localized to generate any significant responses. Figure 4
(Possible braided stream/alluvial fan deposit in seismic designated by dashed line) is a probable
example of the overall lenslike geometry and poor internal seismic character of such a deposit.

6c. Meandering Stream Channels

With greater distance from the sediment source area, a meandering river becomes typical (
Figure 1 , idealized block diagram showing meandering river system over region of low slope and
continual subsidence). Alluvial flood plains cut by a single meander channel occur in regions
characterized by relatively low gradients, higher suspended load component, fine- to medium-
grained sediment, and more continuous (nonseasonal) discharge. Sand bodies are created as
point bar sands resulting from channel migration.

Summary of Facies Characteristics

Lithology

• overall, approximately 1:1 sand/shale ratio

• point bar: flood-plain siltstone/shale, medium to fine sand, well-sorted and


channel pebble lag

• abandoned channel: oxbow lake siltstone/shale channel lag

• flood-plain shale, coal

Sedimentary Structures
• scour and fill

• surface exposure features; mud cracks, raindrop impressions


• ripples

• planar bedding

• trough, tabular crossbedding oriented pebbles, current lineations

Paleontology
• potentially diverse: vertebrates, plant remains, nonmarine mollusks, gastopod shells,
spores, pollen, burrows, footprints
Geometry
• point bars: stacked to relatively isolated lenticular sand bodies

• channels: continuous and discontinuous "shoestrings," sometimes encased in


less permeable sands/silts or flood-plain shales

Associated Facies
• most common: deltaic, shoreline/marine shelf, lakes, braided streams
Diagnostic Evidence

Cores
Core sampling of point bar sands should show the overall fining-upward sequence of sedimentary
types and structures illustrated in Figure 2 (Idealized "outcrop" showing upward succession of
grain size and sedimentary structures in preserved point bar). Such sequences are often
truncated by overlying channels and the entire suite may not be seen.

Logs

Figure 3 (Well log showing two upward-fining point-bar sand bodies. Note the characteristic
bell-shaped curve for channels. There is little, however, to distinguish these filled channels as
alluvial.) presents several logs that show the variations and relationships in meandering stream,
alluvial floodplain sediments. Two point-bar sequences are in evidence.

Both are surrounded by overbank flood-plain shales. Note how the gamma ray curve shows the
abrupt change from shale to sand at the base of each channel, as well as the fining-upward, bell-
shaped curve as point-bar sand grades into flood-plain shale at the top of each channel
sequence.
The dipmeter log for such a section will be a bit complex, but will show three main depositional
surfaces ( Figure 4 , Idealized dip log showing both the filled-in red motif (left) and the upward-
increasing blue motif, which indicates individual crossbed sets. Note that the blue pattern
(right) depends upon a narrow dip correlation interval (usually less than 10 ft.), so that both toeset
and foreset dips can be recorded by the logging tool): structural dip (green motif), major accretion
slopes (red motif), and crossbedding (blue motif).

Seismic
In the subsurface, channels generally create abrupt changes in lithology. Their seismic "visibility"
should, therefore, be pronounced. At the same time, where the sharp erosional base and sides of
the average channel make for good velocity contrast, the upper part of the average channel
grades into flood-plain deposits, and thus will not generate high-quality reflections. As a result,
the typical lens shape of most channels should be only relatively clear on high resolution seismic
lines, as shown in Figure 5 . (Seismic expression of a river-cut channel. Note the abrupt
termination of flat-lying reflections against the channel flanks and the change in seismic character
between these reflections and those within the channel. Note also the steeper slope of the right
flank of the channel, possibly indicating that this was the cut bank.)

6d. Eolian Dunes


The bed forms into which sand settles when transported by wind are mainly asymmetric ripples
and dunes whose overall geometry is much like that of their subaqueous counterparts. Most
dunes preserved in the sedimentary record appear to be the transverse type ( Figure 1 , Cross
section of barchan or transverse dune showing the various bedforms and slipface surface).
The dynamics of eolian and aqueous movement are basically similar: they both involve
granular solids being moved by and within "fluids." This is probably the main reason why the
eolian environment is particularly difficult to distinguish in the subsurface.

Summary of Facies Characteristics

Lithology

• clean, well-sorted quartz sandstones (orthoquartzite)

• scattered, local interdune shale, evaporite, or lag lenses

• layers of heavy mineral concentrations

• pure carbonate sand, much more rare

Sedimentary Structures
• primarily large- to giant-scale crossbedding with high angle foresets (20º-35º)

• surface exposure features (rain drops, rootlets, tracks and trails, etc.) in
interdune lithologies

Paleontology
• rare vertebrate remains

• oxidized spores, pollen

Geometry
• usually sheetlike, upper surface often planed by transgressive seas
Associated Facies
• potentially variable: alluvial fans, braided streams, sabkha, playa in interior arid
basins; barrier island, lagoonal and shallow shelf facies in coastal settings, often
complexly interbedded with water-laid deposits

Diagnostic Evidence

Cores

Samples of eolianite sections are commonly composed almost entirely of clean, well-sorted
quartz sandstone (often called orthoquartzite or quartzarenite) ( Figure 2 , Idealized vertical
sequence of eolian dune and interdune sediments). Detailed sedimentological analysis has
not proven its unqualified worth in strictly distinguishing dunes from some transitional marine
facies. Since eolian sand is often reworked from older deposits, such study may reveal mostly
"inherited" features. Thick sets of monotonously consistent crossbedding are the prominent
sedimentary structures found in most cores. Traces of oxidized "impurities" between sand grains
— whether ferric iron, spores, or heavy minerals — can be significantly diagnostic.

Logs
Figure 3 (Log motifs for eolian sands. Note well-developed blue pattern of upward-increasing
dips along the toeset-foreset transition in individual dune units ) shows a suite of logs typical of
the eolian Rotliegendes (Permian) group, a productive reservoir in the North Sea. Despite an
overall blocky appearance, the gamma ray curve can be divided into approximately 50-ft
increments, all bordered by narrow spikes of higher radioactivity. Thus, the general profile can be
more accurately described as "saw-toothed." Each of the small kicks (which are more obvious on
the density log curve) is asymmetric, with a gentle upward decrease in gamma ray API units.
They are caused by the finergrained, mica-containing layers of each new dune that abruptly
truncate the foresets of the underlying dune unit. The 50-ft interval is also strongly evident on the
dipmeter. Each increment begins at the base with low-angle dips (toeset beds), which then
increase upward until reaching a maximum of about 25º to 35º (foreset beds). This maximum is
the most conspicuous part of the dipmeter log and indicates both the large size and consistent
orientation of the crossbedding. Dip azimuths are very constant, directly indicating the downwind
direction. This, in turn, reveals the local elongation of the sand body transverse (perpendicular) to
wind direction.

Seismic

In general, subsurface dune deposits are not detectable as such by existing seismic methods.
Sheetlike geometry, association with unconformities, and absence of good internal reflectors are,
as mentioned, also typical of the overall response generated by braided stream sediments which
may over- or underlie eolianites and thus further mask them. Seismic data, therefore, are perhaps
most useful in delineating the depositional limits, rather than the actual lithology, of a potential
dune reservoir.

6e. Lacustrine Deposits


Unlike the previous environments we have looked at, lakes usually do not define a single facies,
but a collection, and might better be considered to represent a facies group ( Figure 1 , Block
diagram illustrating the major facies and subfacies of Lake Unita, northeastern Utah, as it is
interpreted to have looked in the Eocene. Alluvial, marginal-lacustrine, and open-lacustrine
depositional environments existed simultaneously). Lakes that have occupied intracratonic basins
can, to some extent, be considered as small inland seas in terms of their major facies. They may
be bordered by coastal alluvial plains with swamps, lagoons, and barrier islands (tidal flats are
notably absent). They may also be the site for deltas, which form at major river mouths, and from
which turbidity (subaqueous gravity slide) currents transport sediment into the basin center,
creating subaqueous fan deposits. This means that lithology is often completely undiagnostic for
this environment.

The far more subdued water turbulence of the lacustrine environment — waves, longshore and
subsurface currents — as well as its different geochemistry, sometimes effects significant,
partially diagnostic differences from marine counterparts. Lacustrine sediments, for example, are
often much more finely bedded (laminated) and contain better preserved plant debris than those
in most marine settings (lagoons being a major exception). Certainly, paleontology is the most
prognostic indicator, but experience dictates that reworking and redeposition of nonmarine fossils
in marine facies is common.

In general, the consistent indication of aqueous deposition and nonmarine fossils, as well as the
"negative evidence" offered by the lack of marine biota, together indicate the probability that the
facies under consideration is lacustrine. More broadly, tectonic settings can also afford a strong
clue. Small continental basins, as well as rift graben-type basins associated with continental
breakup, are strong candidates for having at one time or another played host to large lakes.

6f. Deltas
The delta environment contains diverse settings for sandstone deposition (Figure1,Sand deposits
of a delta system). In the upper delta plain, point-bar or braided-stream channel sands may be
deposited. When streams contain high sediment bedload or when marine processes dominate
(high-energy deltas), these alluvial channel sand deposits may extend over the entire delta plain
to the shoreline. In river-dominated deltas of low marine energy, alluvial channel deposits of the
upper delta plain give way, through stream bifurcation, to a network of essentially straight
distributary channel deposits on the lower delta plain. Surrounding these channels are fine-
grained bay-fill sediments, often containing coarsening-upward sandy sequences deposited by
crevasse subdeltas.

The subaqueous delta contains distributary front bar sands that may be reworked into barrier
islands by marine processes in abandoned portions of low-energy, river-dominated deltas. In
high-energy deltas, winnowing of fine-grained material by waves, currents, and tides creates a
variety of sand deposits along the shoreline, in the form of barrier islands, tidal channels, and tidal
sand sheets.

The characteristics and diagnostic evidence of braided stream and point bar sands deposited in
the delta environment are essentially the same. Crevasse subdelta sands generally form minor
petroleum reservoirs. We shall touch on the two major deltaic sandstone facies: distributary
channel sands and distributary mouth bar sands.

6f.1. Facies Characteristics of Distributary Channel Sands

Lithology

• fine- to medium-grained sandstone, moderate- to well-sorted

• fining-upward grain-size profiles

Sedimentary Structures
• contorted bedding

• ripple formations
• planar bedding

• trough, tabular crossbedding

• scour base

Paleontology
• burrows

• organic plant debris

• faunal remains usually absent

Geometry
• linear, straight to sinuous

• 10 in to 30 in thick

• 1 km to 5 km wide

Associated Facies
• fluvial meander point bar or braided stream

• interdistributary bay, crevasse subdelta

• distributary mouth bar

Diagnostic Evidence for Distributary Channel Sands

Cores and cutting samples should show a suite of lithologies and structures similar to that shown
in Figure 2 (Idealized lithogenetic sequence of vertically stacked point bars from upper delta plain
area). An upward-fining sequence of medium- to fine-grained, moderately to well-sorted
sandstone is typical. Sedimentary structures vary from large-scale cross strata in lower portions
of units to interbedded ripple cross laminations and planar lamination in upper parts. Fragments
of plant and coaly material are common.

Logs

SP/gamma ray curves typically display blocky to upward-fining "bell" shapes with abrupt bases.
Curves are often jagged, reflecting shale laminations within the sand, and dip-meters in
distributary channel sands tend to display red "slope" patterns of increasing dip with depth (
Figure 3 , Sp and dipmeter logs of a distributary channel sand reservoir, offshore Louisiana, with
a schematic cross section showing location of logs within the channel. Note red pattern dip
azimuths point toward channel axis). These shapes reflect deposition on lateral accretion
surfaces and dip azimuths usually point toward the channel axis and, thus, are normal to channel
strike.

6f.2. Facies Characteristics of Distributary Mouth Bar Sands


Lithology

• in proximal bar: clean, well-sorted coarse- to medium-grained sandstone

• in distal bar: coarsening-upward sequence of fine sand, silt, and clay

Sedimentary Structures
• in proximal bar: small-scale cross laminae and current ripples

• in distal bar: small-scale cross laminae, small scour and fill, and graded sand
units

Paleontology
• abundant microfossils in prodelta clays at base of sequence with minor
bioturbation

• microfossils and bioturbations decrease upward

• small burrows and shell remains in distal bar

• laminations of organic debris in upper sand body (proximal bar)

Geometry
• elongate in seaward direction with high river influence; arcuate to cuspate-
shaped, with increased wave and marine current action

• up to 130 in thick and 10 in wide

Associated Facies
• prodelta marine shale

• delta plain and interdistributary bay silts and clays

• distributary channel sands

• crevasse subdelta silts

Diagnostic Evidence for Distributary Mouth Bar Sands

Cores should typically show lithologies and sedimentary structures illustrated in Figure 4
(Lithologic column of distributary mouth bar deposit ) i.e., distal shales and silts coarsening
upward to coarse-to-medium, well-sorted sand in upper bar. Sedimentary structures are
primarily ripple laminations in fine- to medium-grained sandstones.
Logs

The electric log in Figure 5 (Gamma-ray/SP and dipmeter log of distributary mouth bar
sequences, subsurface Gulf of Mexico. Note blue current dip motifs pointing in direction of
current flow ) shows the SP/gamma ray curve of distributary mouth bars displays an overall
funnel-shaped, coarsening-upward profile. An abrupt break is usually seen at the top of the
curves, reflecting the sharp change from clean, well-sorted sand of the uppermost bar to a
capping by fine-grained sediments.

On dipmeters, distributary mouth bars are often characterized by patterns of upward-increasing


dips (blue patterns) ( Figure 5 ). This pattern reflects deposition by progressively stronger
currents as a bar is built up into shallower water. Dip azimuths generally point in the direction of
current flow (seaward) , but variations may be considerable.

Seismic

Oblique progradation is the type of reflection configuration typically associated with fluvial delta
systems. Sediment input in this environment is high compared to sea level rise/basin subsidence,
resulting in significantly more lateral progradation than vertical aggradation. The oblique
configuration is distinguished by reflections that terminate by toplap at or near the upper surface,
and by downlap at the base ( Figure 6 , Oblique-progradational seismic reflection pattern typical
of deltaic systems).

An actual map view of an ancient deltaic channel or bar sand may be revealed by a horizontal
slice through a block of 3-D seismic data. Figure 7 (Horizontal slice through block of 3-d seismic
data from Gulf of Mexico, showing lenticular-shaped distributary channel sand. Superimposed
structural contours show brightest (darkest) portion of channel, where gas is indicated, is
structurally high) is such a horizontal section from the Gulf of Mexico displaying variations in
reflection amplitude along a structurally interpreted horizon (Brown 1985). We can clearly see a
bifurcating distributary channel delineated by a zone of high amplitudes (darkest tone) cutting
from northeast to southwest across the section.
6g. Fan Deltas

Fan deltas are alluvial fans that prograde out into a standing body of water from an adjacent
highland (Holmes 1965). As such, they generally develop on the flanks of basins next to fault-
bounded, elevated source areas (Figure 1 , Typical fan-delta tectonic setting on flank of rift
valley). When fan deltas form adjacent to contemporaneous faults, thick wedges of coarse-
grained deposits accumulate.
Fan deltas have only recently been recognized as important oil and gas reservoirs (Ethridge and
Wescott 1984). Rapid facies changes and association with tectonically active basin margins
create favorable stratigraphic and structural trapping conditions. Furthermore, potential reservoir
beds are often in close juxtaposition with marine hydrocarbon source rocks.

The cross sections in Figure 2 (Idealized vertical sequence), Figure 3 (paleogeographic


reconstruction), and Figure 4 (cross section of shelf-type fan deltas based on data from U.S.
midcontinent Pennsylvanian-Permian granite wash studies. Based on studies by McGowen,
1970) illustrate the distribution of facies within a shelf-type fan delta. This type of fan delta forms
on the broad shelves that typically border intracratonic and plate-divergent basins and often
develops extensive progradational sequences.

We see that the proximal and medial parts of the fan, collectively called the fan plain, occupy the
exposed portion of the fan-delta system. The distal fan and prodelta environments constitute the
subaqueous portion of the fan system.
Summary of Facies Characteristics

Lithology

• fan plain: poorly sorted, coarse-grained, sands and gravels; often highly arkosic

• distal fan: well-sorted, coarsening-upward sequences of sand and gravel,


grading offshore into prodelta shales and possible marine limestones

Sedimentary Structures

• crude to unbedded (fanglomerate) in proximal fan

• large-scale, tabular, and trough crossbedding in braided channels in medial fan


with occasional horizontal stratification

• parallel-laminated to massive delta-front sands

Paleontology

• rare vertebrate bones, plant debris in fan plain

• shell fragments in delta-front sands

• microfossils (marine or fresh water) in prodelta shales

Geometry

• overall fan-shaped in plan view

• wedge-shaped in radial profile

• convex-upward in transverse profile

• subaqueous distal facies elongate in seaward direction in fluvial-dominated


fans; arcuate to cuspate-shaped, with increased wave and marine current action

Associated Facies

• fault-generated mountain fronts, proximally

• marine or lacustrine shales and limestones, distally

Diagnostic Evidence

Cores and cuttings should show a high ratio of coarse- to fine-grained sediment, often highly
arkosic, and an overall coarsening-upward succession in vertical sequence ( Figure 5 and
Figure 6 , Hypothetical vertical sequence in shelf-type fan delta based on studies of fans and
fan deltas along the southern Alaska coast).
Logs

Figure 7 (SP and resistivity log of the fan-delta Ivishak formation, main reservoir in the
Prudhoe Bay field, Alaska ) a log of the reservoir Ivishak formation, Prudhoe Bay field, Alaska,
illustrates the overall coarsening-upward Sp profile of a fan delta sequence. The base of the
sequence consists of very fine-grained sandstones and mudstones of the offshore that coarsen
upward into fine-grained, well-sorted sandstones deposited in a beach-bar shoreline complex.
Overlying the shoreline sands are coarse-grained sandstones deposited in braided streams of the
distal fan plain that are capped by conglomerates of the proximal alluvial fan. In this example, a
sequence of braided stream channel sand ("upper sandstone sequence") from a following cycle
overlies the proximal conglomerate facies.
Seismic

In a study of seismic reflection patterns from offshore Brazilian basins, Brown and Fisher (1977)
presented patterns characteristic of fan delta/shelf facies. They found the reflection patterns
developed in response to proximal, medial fan facies to be poorly defined and parallel-layered to
reflection free ( Figure 8 , Seismic facies patterns characteristic of fan delta/shelf reflections,
generalized from offshore Brazil seismic section). Reflection continuity is very poor to absent,
and the external geometry of the reflection units is wedge-shaped, thickening toward the source
area or toward bounding basement faults. The distal fan and prodelta facies contain some poorly
defined, inclined to horizontal, slightly divergent, layered reflectors increasing in number
basinward and grading into well-developed shelf reflections.
6h. Coastal Barrier Islands

Coastal barrier sand bodies are generally narrow, wave-built, sandy islands or peninsulas that
form parallel to shore ( Figure 1 , Figure 2 and Figure 3 , Idealized block diagram and
cross sections showing principal environments and facies of a regressive barrier island system).
As topographic features, they can be perennially emergent or exposed only during periods of low
tide. They most frequently occur as a linear trend of individual islands separated by tidal inlet
channels. The main sand body of a barrier island is created almost entirely by relatively high-
energy, shallow marine processes. In many instances, subaerial reworking by onshore winds
leads to the formation of a capping dune field.

The offshore-to-beach profile of Figure 4 (Typical fan-delta tectonic setting on flank of rift valley )
shows the progression of specific depositional zones. The offshore or shelf zone grades
landward into the lower and then upper shore face zones, which form the seaward portion of the
barrier island. Above the mean low water level is the beach/ dune zone.

Summary of Facies Characteristics

Lithology

• lower shoreface: fine- to medium-grained sand

• upper shoreface: medium- to coarse-grained, well-sorted sand

• beach: medium- to coarse-grained, well-sorted sand, occasionally with


conglomerate

Sedimentary Structures
• lower shoreface: small-scale cross lamination and parallel stratification, often
hummocky; abundant bioturbation

• upper shoreface: high-angle trough cross stratification, planar tabular bedding

• beach: low-angle, planar stratification, dipping seaward; possible high-angle


cross stratification dipping landward
Paleontology
• macrofossils (bivalves and gastropods) and shell fragments

• trace fossils: straight burrows of low-level suspension feeders in subaqueous


barrier

• rootlet horizons (uppermost beach)

Geometry
• thickness: 10 in (low-energy coasts) to 30 in (high-energy coasts)

• elongate: 20 kin to 100 km in length on microtidal (0-2 in) coasts

• stunted ("drumstick"-shaped) 3 km to 20 km in length on mesotidal (2-4 m)


coasts)

• barriers generally absent on macrotidal (>4 in) coasts

• isolated, shoestring bodies when formed by rapid transgressing seas

• overlapping series of bars when formed by regressing seas

Associated Facies
• marine shelf shales

• lagoonal silts and shales

• tidal channel, tidal delta/inlet and washover fan sands

Diagnostic Evidence

Cores

As illustrated by Figure 5 (Theoretical vertical sequence of a barrier island system based on


studies of modern deposits on Oregon coast), cores and cuttings should reflect the following
basic characteristics of a regressive barrier island sand sequence:

• a progressive and fairly regular upward increase in grain size from silt/clay to
coarse sand and possibly conglomerate, with maximum grain size usually
occurring in the upper shoreface.

• a simultaneous upward improvement in sorting, from fair to good in the lower


shoreface, to excellent within the upper shore face and beach.

• a general upward increase in both the abundance and scale of cross


stratification, indicative of higher energy levels.

• a general upward decrease in the disturbance of primary stratification due to


bioturbation.
Logs

SP and gamma ray logs through barrier island sands commonly display the smooth funnel shape
that reflects a regular upward increase in grain size, sorting, and permeability. Greater amounts
of fine-grained material depress and round off this curve, while barrier island sands that are
almost entirely free of clay and silt generate a blockier profile. Figure 6 (Generalized electric log
patterns across a barrier island system, showing changes in log shape depending on location
and relative richness of sand versus shale) shows how log curves ideally vary according to
changes in the amount of fine and coarse material and to location within the barrier system.
As shown in Figure 7 (Gamma ray log and dipmeter motifs for barrier island sand bodies),
dipmeter patterns for barrier bar sands usually display an upward-increasing blue motif reflecting
the concave profile of the seaward depositional slope. Dips within the barrier sand body may,
when plotted on a rose diagram, reveal a bimodal pattern. The lower angle dips, which define the
main blue motif, represent seaward-inclined beds formed by wave swash, while higher dips with
opposite azimuths reflect landward-dipping foresets, presumably from ridge and runnel
deposition.

Seismic

A typically seismic response should generate a high amplitude reflection from the sharp upper
contact between the coarser beach/dune or upper shore face sands and the overlying marine or
lagoonal shales. A sharp but diminishing reflection is generated from the sides of the sand body,
caused by the downward-fining in grain size and a weaker response marking the transition to the
fine-grained base of the sand body.
In a profile showing three pulses of barrier island regression (part a of Figure 8 , Seismic profile
showing three pulses of barrier-bar regression), note the high amplitude reflection caused by the
contrast between the upper barrier sands and overlying lagoon/marsh material. The general
depositional slope and the direction of progradation are to the right, as modeled in the
accompanying cross section (part b of Figure 8 , Block diagram showing how transgressive-
regressive sand bodies are composed of a stair-step multitude of individual bars).

6i. Continental Shelf Sands

Shelf sands form as linear ridges usually oriented oblique to the shoreline, or as sheetlike
deposits. They occur between the lower shoreface and shelf edge of continental shelves ( Figure
1 , Occurrence of sand deposits on the continental shelf) and in broad, relatively shallow
epicontinental seas, such as the North Sea. Tidal- and storm-generated currents have been
shown to be the two most significant agents responsible for shelf sand deposition.

Summary of Facies Characteristics

Lithology

• fine- to coarse-grained sand, moderately sorted, possible pebble conglomerate at top of


unit

• generally coarsening-upward grain size profile

• minor to abundant glauconite

• occasional shale laminations and shale clasts

Sedimentary Structures
• predominately moderate angle trough and planar crossbedding

• some planar laminated bedding

• ripple stratification in lower units, often hummocky

• bioturbated in lower units

• possible scour at base of some high-energy deposits


Paleontology
• marine shelf foraminiferal assemblages in associated finegrained rocks

• macrofossil shell "hash" at scour base of some high-energy ridges

• Cruziana and Zoophycus ichnofacies

Geometry
• commonly series of parallel ridges, asymmetrical in cross section, up to 50 km long, 3
km wide, and 40 in thick

• less commonly sheetlike, up to 20,000 sq kin in area and up to 12 in thick

Associated Facies
• surrounded by marine shelf shales

• possible lower-shoreface fine sands and silts laterally shoreward of some shelf
ridges

Diagnostic Evidence

Cores
The upward-coarsening lithofacies sequence, illustrated in Figure 2 (Idealized lithologic sequence
of the Viking formation, Joffre-Joarcam area, Canada ), from the productive Cretaceous Viking
formation of Alberta, Canada, is characteristic of many shelf sand ridge deposits. The basal
facies consists of a burrowed, silty gray shale. This facies is overlain transitionally by ripple-
bedded sandstone intercalated with silty shale and containing abundant burrowing.
Interchangeable with this ripple-bedded sandstone facies is a bioturbated, shaly, fine-grained
sandstone.

Next in vertical sequence is a trough crossbedded, fine- to very coarse-grained, well- to


moderately sorted sandstone. Shale clasts are common and the sandstone contains abundant
glauconite. Generally this facies has a sharp lower contact and a gradational upper contact. A
pebble conglomerate occasionally forms the top of the sequence.

Overlying the sequence may be another interval of bioturbated or rippled sandstone and shale,
which, in turn, is overlain by crossbedded sandstone.

Logs

SP/gamma ray log profiles may show a variety of shapes: funnel (coarsening-upward), blocky,
serrated, and more rarely, bell-shaped (fining-upward). The type of profile depends on the
amount and occurrence of dispersed clay and clay intervals, which, in turn, are dependent on the
nature of shelf near-bottom currents.

Therefore, log shapes of shelf sands tend to reflect flow regimes. In general, a funnel-shaped,
coarsening-upward profile (the most common of shelf sand log profiles) suggests a storm/wave-
dominated shelf. A blunt-base, blunt-top signature is more characteristic of tidal-current sand
bodies (Selley 1976).
Figure 3 (Gamma-ray neutron-density log of the Cretaceous Shannon formation, oil reservoir of
Hartzog Draw field, Wyoming, showing coarsening-upward grain-size profile and
corresponding upward increase in porosity typical of a shelf sand ) from the Shannon sandstone
of Wyoming illustrates the coarsening-upward, funnel-shaped profile common for storm-emplaced
shelf sands. Figure 4 (Log of the Cretaceous sub-Clarksville. a shelf sandstone in Iola field,
Texas, with plots of texture and composition, showing a fining-upward sequence probably
resulting from rapid deposition by waning storm currents) shows a log from the Cretaceous sub-
Clarksville sandstone of Texas. Here however, the log and grain-size plot show a fining-upward
sequence characteristic of rapid deposition by waning current flows, probably from geostrophic
storm currents.

Figure 5 (Gamma-ray log of shelf tidal current sand body from the North Sea, showing
characteristic blocky shape with blunt base and top associated with many tidal sands ) is a
gamma-ray log from an undisclosed North Sea location where the sand body was postulated to
originate on a tide-dominated shelf. The log profile has the characteristic blocky shape with blunt
base and top associated with many tidal sands.
Seismic
Shelf sands usually coarsen upward from a shale base to a coarse sand or conglomeratic top that
is abruptly overlain by marine shale. Laterally these bodies are fringed with tight silt that grades
into marine shale. The seismic model in Figure 6 (Seismic model of a thin shelf sandstone,
Cardium formation, Alberta, Canada ) reflects these overall lithologic changes by showing a
strong event at the upper sharp contact and a lower-amplitude event at the gradational base.
As the reservoir becomes transitional into silt, updip to the right, there is a gradual decrease in
amplitude and the exact boundary between porous reservoir rock and tight silt is difficult to
determine seismically.

6j. Deep Sea Sands


Deep sea research of present-day ocean bottoms, along with petroleum exploration in ancient
basins, has shown that a particular type of deep water sedimentary facies is characterized by
thick sequences of laterally extensive interbedded sands and shales. These deposits have been
variously called deep sea sands, deep water fans, turbidites, submarine fans, and turbidite fans.
Although fan-type deposits make up the bulk of the sediment, feeder channel sands and slump
deposits can be important subfacies ( Figure 1 , Model of a submarine fan). And, though turbidite
currents are believed to be the primary depositional process, processes like debris flow and grain
flow can be significant in the proximal fan area.

Summary of Facies Characteristics

Lithology

• pebbly conglomerate and massive sand in channels


• upward-fining, vertically graded turbidites that constitute overall coarsening-
upward sequences of sand, silt, and shale in mid- to lower-fan and inter-channel
areas

Sedimentary Structures
• scoured erosion surfaces

• dish structures and pillars in channel sands

• laminated sands, cross-laminated sands, and laminated, often convoluted, silts


and fine sands in turbidites

Paleontology
• macrofossils (in situ) rare

• micro fossils common in finer-grained sediments

Geometry
• fans are mound shaped, concave downward in strike profile

• fans are lenslike, concave upward in dip profile

• individual channels long and narrow or coalesced into sheets

Associated Facies
• marine pelagic shales

• slope shales

Diagnostic Evidence

Cores

A diverse range of sediment from boulder beds to fine silt and clay is characteristic of deep sea
fans. Figure 2 and Figure 3 (Hypothetical stratigraphic sequence of a prograding submarine
fan: C.T., classical turbidite; M.S., massive sandstone; P.S., pebbly sandstone; D.F., debris flow;
S.L., slumps; C.G.L., conglomerate. Arrows show thickening-upward and thinning-upward
sequences) shows the overall stratigraphic sequence typically developed by a prograding fan.
The lower portion of the sequence consists of often incomplete turbidite sequences (CT). The
upper portion of the sequence is dominated by cut-and-fill channel sediments composed of
massive sand (MS), pebbly sand (PS), and conglomerate (CGL). The main feeder channel in the
upper fan may be filled with debris flow (DF) sediment characterized by massive, poorly sorted
sand with clasts of coarse gravel.
Figure 4 (Idealized turbidite sequence showing Bouma subdivisions Ta throught Te with
hemipelagic subdivisions for the Te unit) shows the ideal lithologies and sedimentary structures
anticipated in cores of the turbidite units. With increased distance away from the source the
coarser, lower units of the sequence become missing from nondeposition. Thus, in the lower fan,
only the upper, finegrained portions of the sequence are deposited. Glauconite and
carbonaceous detritus are often found mixed together if sediment is derived from both marine and
deltaic sources.
Logs

Figure 5 (Blocky to fining-upward gamma-ray and dipmeter motif of submarine feeder channels)
shows the typical blocky or fining-upward SP/gamma ray profiles of feeder channels of the upper
fan with random, scattered dips displayed by the dipmeter.

Figure 6 (Gamma-ray profiles of proximal fan sands showing fining-upward channel sands,
and dipmeter showing red "slope" motifs dipping into the center of the channels) shows the
thinner fining-upward SP/gamma ray profiles developed in channels of the proximal (mid) fan
area. Dipmeters may display red "slope" motifs dipping into the center of the channels in a
direction perpendicular to channel axes. Thin blue current patterns are often absent because
crossbedding is usually not well developed in deep sea sands.

Figure 7 (Gamma log profiles of distal fan sands showing coarsening-upward progradational
sequences, and dipmeter showing blue dip patterns pointing in the direction of fan
progradation) shows the upward-coarsening SP/gamma ray profiles of the distal (lower) fan
progradational sequences. These larger sequences in turn are made up of individual upward-
fining turbidite units. Combined with the presence of marine pelagic shale intervals, the resulting
SP/gamma ray profile of a distal fan displays a "nervous" back-and-forth character. Blue dip
patterns that may be evident point in the direction of fan progradation.

Seismic
Perhaps the most direct seismic indicator of submarine fans is a mound-shaped seismic
sequence with an internal hummocky or chaotic reflection pattern.

Figure 8 (Seismic section across the Frigg field, a giant gas field in the North Sea. The
pronounced mound with hummocky reflections from 1.8 to 2.0 sec is a submarine fan, and the
"flat spot" around 2.0 sec is a gas-liquid contact ) is a seismic section across the Frigg field, a
giant gas accumulation in the North Sea, which produces from a submarine fan. Note the
pronounced mound with hummocky reflections from 1.8 to 2.0 seconds, centered under shotpoint
150. The high amplitude reflection at 2.0 seconds is a "flat spot" representing a seismic reflection
off the gas-liquid contact.

The presence of canyons or troughs on a seismic section may indicate the presence of a
submarine fan located basinward of these features. Submarine fans also may be present beneath
or basinward of features displaying clinoform (sigmoid or oblique) patterns.
Case Study: Marchand "C" Sandstone, Anadarko Basin, Oklahoma
A typical environmental study of a productive sandstone was described by Baker (1979).

The Pennsylvanian (Upper Carboniferous) Marchand "C" Sandstone of the Anadarko Basin,
Oklahoma produces oil from over 90 wells in the East Binger field ( Figure 1 , Location map of
East Binger field, Anadarko Basin, Caddo Co., Oklahoma). Regional structure consists of a
monoclinal surface with local structural noses and saddles; hydrocarbon entrapment is purely
stratigraphic with porous, permeable sands grading laterally into siltstones and shales.

The first step in the study was to determine the external geometry of the sand body by
constructing an isopach map of the gross clean sandstone ( Figure 2 , Isopach of gross sand,
Marchand "C"). The sand body was shown to be an elongate, northwest-southeast trending
sand ribbon extending over 8 mi in length and 1 to 1.5 mi in width. Sand thickness ranged up to
80 ft, and the sand body was asymmetrical in shape with a steep southwest flank.
A northeast-southwest stratigraphic cross section was then constructed in a direction normal to
the long axis of the field. With streaks of "hot-shale" and local marker beds providing correlations,
the section showed the sandstone to thicken at the expense of the underlying shale, a classic
indication of a channel deposit ( Figure 3 , Stratigraphic cross section, C-C', across East Binger
field, showing sandstone thickening at expense of underlying shale). However, based on
regional isopach mapping ( Figure 4 ) the long axis of the sand body apparently parallels
depositional strike. Channel deposits usually extend approximately normal to depositional strike.
Thus a channel interpretation seemed unlikely.
Thin sections and hand samples of the sand revealed no woody or carbonaceous detritus,
thus indicating a turbulent environment (beach or shoal). In contrast, glauconite, fecal pellets, and
crinoid fossils were present in the sand, indicating a marine origin. Combining this information
with the blocky gamma ray profiles and applying it to Selley's model a sub-tidal sand ridge was
suggested. ( Figure 5 , Four characteristic gamma log motifs. From left to right: thinly
interbedded sand and shale; an upward- coarsening profile with an abrupt upper sand-shale
contact; a uniform sand with abrupt upper and lower contacts; and, furthest right, an upward-
fining sand-shale sequence with an abrupt base. None of these motifs is environmentally
diagnostic on its own. Coupled with data on their glauconite and carbonaceous detritus content,
however, they define the origin of many sand bodies.)

However, the stratigraphic cross section showed the sand body to be emplaced into the
underlying shale; thus it was unlikely to have been formed as a "ridge."

A core study revealed underlying burrowed, fossiliferous marine shale scoured into by a basal lag
deposit, ripple and megaripple crossbeds, herringbone cross-stratification, shale lamina, and
slumping of sediments - all very similar to sequences seen in present-day subtidal sand ribbons
that are formed by tide-generated bottom currents in the North Sea.

It was finally concluded that the Marchand "C" Sandstone was indeed probably formed as a
subtidal sand ribbon but infilled a scoured depression similar to the depressions present on many
modern tide-dominated shelves.
7. CARBONATE RESERVOIR FACIES

General Considerations
The essence of successful exploration in carbonate rocks is the discovery of porosity. This is so
because porosity development in carbonate rocks is generally much more variable and erratic
than in terrigenous sandstones. Diagenesis is more important as an agent that preserves,
creates, and rearranges porosity. There are, moreover, many more types of porosity in
carbonates than in sandstones, even though carbonates are, on the whole, less porous than
sandstones.

Primary porosity refers to pore space that has been preserved since the sedimentary rock was
deposited. Most commonly, it includes void spaces between sedimentary particles, as well as
original cavities within certain skeletal grains.

Secondary porosity refers to pore space that has been created after deposition by diagenesis or
fracturing. One important aspect of diagenesis involves solution and removal of sedimentary grain
material or carbonate cement. Usually, such porosity can be seen to "cut across" grains and/or
cement, although even secondary porosity development often shows a strong preference to be
guided by original depositional fabric and primary porosity occurrence.

Though carbonate diagenetic processes are variable, complex, and difficult to predict, there is
often a close, early connection between the original depositional environment and subsequent
diagenetic changes. Carbonate explorers have thus learned to focus their attention on
depositional models and facies trends (within the context of a stratigraphic framework) as the best
guide to prediction of reservoir rocks.

Lithologic Parameters of Carbonate Facies


The fabric of modern carbonate sediments consists of three end-members (Folk 1962):

• lime mud or micrite — mostly clay-size particles of aragonite and fine, silt-size
particles of skeletal debris.

• grains (allochems) — gravel, sand, and silt-size carbonate particles that


originate within the basin of deposition.

• pore space — openings that commonly become filled by clear, sparry calcite
cement.

Most carbonate grains are biogenic in origin and produced locally. Thus, grain types are strong indicators
of depositional environment. The shape, size, and sorting of carbonate sedimentary grains relate both to
environment of deposition (particularly skeletal debris) and to transporting currents.

Carbonate Grain Types

There are five main types of carbonate sand grains: skeletal grains, ooids and pisoliths,
pelletoids, aggregates and lumps, and lithoclasts.
Skeletal grains range from unabraded whole shells or colonial masses to well-rounded, clastic
shell fragments. Skeletal grains may be divided into several types, depending on the type of
organisms producing the grain (Table 1, below). Figure 1 (Ginsburg 1956) demonstrates
consistent patterns of variation in sediment grain size and constituent composition of skeletal
grains in studies of south Florida sediments. These distributions have been shown to be
related to submarine topography, areal geography and hydrography. Analogous lateral
successions are noted in ancient carbonate successions as well, providing powerful evidence of
depositional environments.

Table 1:

1. Sheathed and spiculed skeletons are those in which the mineral matter is small (silt-
fine sand size) and loosely held by organic tissue. On death of the organism the organic
tissue decays and the particles are released as fine sediment. Examples: Penicillus,
alcyonarians, corals, sponges, tunicates, holothurians.

2. Segmented skeletons consist of mineral particles linked together by organic tissue.


Death and decay of the organisms most commonly yield sand-size particles to the
sediment. Examples: Halimeda, articulated red algae, echinoderms.

3. Branched skeletons are composed of well-calcified cylindrical or bladelike projections.


The size of fragments found in sediments depends on the original size of the organism,
the size and strength of the branches, and the nature and intensity of organic and
mechanical breakdown to which they are subjected. Examples: some corals (Acropora),
red algae, and bryozoans.

4. Chambered skeletons include all those that are hollow or partly hollow. Chambers
persist after death of the organism, but there are wide variations in the resistance of
different types of chambers to breakdown, depending on their absolute size, wall
thickness, shape, and microstructure. In general, the arcuate shape successfully resists
breakage. Examples: gastropods, serpulid worm tubes, foraminifera, some crustaceans,
pelecypods, some echinoderms, and brachiopods.
5. Encrusted skeletons include all plants and animals that encrust surfaces. The
breakdown of the skeletons in most cases depends mainly on the organic breakdown of
the encrusted surface or substrate. These are mechanically resistent structures.
Examples: some algae, foraminifera, corals, bryozoans, worms, hydrocorallines.

6. Massive skeletons are generally large and hemisperical in shape. They are most
resistent to breakdown because of their size and, in some cases, their microstructure.
Examples: corals and some coralline algae.

Ooids and Pisoliths

Ooids ( Figure 2 , Ooid grains under uncrossed polarizers from Holocene sediment, Great Salt
Lake, Utah. Grain on right shows radiating crystals of aragonite commonly found in
hypersaline lakes. Scale: 1.25cm = 0.10mm) are spherical, smooth, carbonate sand grains,
normally 0.3-1.0 mm in size, consisting of many concentric coatings of aragonite around a detrital
nucleus. They generally form in agitated shallow marine waters and, in time, convert to calcite.
Some ooids may display a fabric of radially arranged fibers, indicating formation in hypersaline
lakes. Carbonate rocks in which the dominant grains are ooids are termed oolites.
Pisoliths (or oncoliths) are coated or laminated sedimentary grains 2-10 mm in diameter ( Figure
3 , Portion of coastal caliche deposit with large pisolites set in laminated crust from Quaternary
sediment, Abu Dhabi. Photograph taken under crossed nicols. Scale: 1.25cm = 0.25mm).
Compared to ooids, pisoliths are larger, more irregular in shape, and are produced in lower-
energy depositional settings. Their laminations are ascribed either to successive algal coatings in
marine and saline lake settings, or to sequential layers of caliche coatings of carbonate particles
in shallow soil profiles.

Pelletoids are round, oblong, or cylindrical carbonate grains with a cryptocrystalline or


microcrystalline texture ( Figure 4 , Pelletoids. (a) Soft, unlithified pelletoids. Reflected light
scale is 1 mm. (b) Hard pelletoids. Reflected light; scale is 1 mm. (c) Hard cylindrical pelletoids.
Reflected light; scale is 1 mm. (d) - (g) Refracted light photomicrographs of pelletoids, showing
the opaque nature of these grains. (d) is under polarized light. Scales are all 250 microns. (h)
micro fabric of an unlithified pelletoid. Note the very small aragonite needles comprising the
matrix, and the relativity large amount of pore space. SEM; scale is 8 microns. (j) Pore space
within a lithified pelletoid. The intragranular cement is composed of large aragonite crystals. SEM;
scale is 25 microns).

Most pelletoids are thought to originate as fecal pellets from various marine invertebrates. Some
may also be severely altered skeletal grains, ooids, or lithoclasts. Modern pelletoids range in size
from 0.1 to 2 mm. Though pelletoids are produced in all environments, they tend to be destroyed
or winnowed out in high-energy settings, and thus are most often preserved in quieter water
environments.
Aggregates and lumps are a class of carbonate grains that include fragments joined together by a
cryptocrystalline matrix. Illing (1954) called such aggregated particles "lumps" and recognized
several types: grapestone ( Figure 5 , An ancient example of a grapestone deposit from the
Lower Cretaceous Cupido Formation, Mexico, taken under uncrossed polarizers. Scale: 1.25
= 0.25), botryoidal lumps, and encrusted lumps. Lithification may be accomplished by organic
binding or submarine cementation. Aggregates and lumps form in sandy shelf areas of moderate
to low wave energy (Milliman 1974).

Lithoclasts are rock fragments resulting from the uplift and erosion of older carbonate rocks or
from the subaqueous or subaerial erosion of contemporaneous carbonate rocks. Lithoclastic
rocks called calclithites consist of detrital grains from older limestones often eroded off rising
mountain fronts. The low abrasion resistance and solubility of carbonate grains mean that such
deposits are usually found in association with major fault scarps and/or very arid climates.

Classification of Carbonate Rock


There are two carbonate classification systems in wide use today. Both systems base
classification on rock texture which, in turn, is controlled essentially by depositional environment.

Folk's classification (1962) of carbonates is based on their three essential ingredients: allochems
(sedimentary grains) , micrite (lime mud), and sparry calcite cement.
Figure 1 (Graphic classification table for limestones ), Figure 2 and Figure 3 (A textural
spectrum for carbonate sediments, showing eight proposed sequential stages. In general,
"low-energy" sediments occur to the left, with successively "higher -energy" sediments to the
right. These stages are quite analogous to the textural maturity sequence in terrigenous rocks.
Because of lack of space, only the biomicrite-biosparite terms are used as representative
examples; for these terms, one can substitute the other allochemical limestone types (for
example, the cell labeled "packed biomicrite" can equally stand for packed intramicrite, packed
oomicrite, and packed pelmicrite). Comparison with the fourfold division used in Folk, 1959 is also
shown )are graphic representations of Folk's basic system. Although Folk's classification and
terminology are widely used, they have several shortcomings; principal among these is that the
application of Folk's terms requires a fairly rigid quantification of proportionate constituents.

Dunham's classification (1962) of carbonate rocks has gained wide acceptance in the petroleum
industry and is now the recommended system. Dunham's system recognizes three basic textural
features that relate to original depositional fabric:

• carbonate mud

• sedimentary grains

• binding by in-situ growth

According to Dunham, a fundamental understanding of depositional environment can be obtained by


distinguishing between sediment deposited in agitated water and sediment deposited in calm water. The
presence or absence of mud, generally a reliable indicator of current strength, should thus be incorporated
into any carbonate classification system. Dunham proceeds to identify the basic carbonate rock fabrics
based on three textural features: abundance of mud, abundance of grains, and signs of binding. ( Figure 4 ,
Dunham's classification of carbonate rocks according to depositional texture.)

• mudstone — composed of micrite with less than 10% grains

• wackestone — more than 10% grains, but grains are mud supported
• packstone — contains some mud; grains are grain supported

• grains tone — no mud content; grains are grain supported, usually well sorted

• boundstone — bound together during deposition; e.g., an ecologic reef

In using Dunham's system, further modifiers as to grain type, grain size, sorting, and carbonate
chemistry should always be used (i.e., medium-grained, well-sorted pellet lime grainstone).

Dunham also recognized a sixth type of carbonate rock that he termed crystalline carbonate — a
rock that contains no recognizable depositional texture.

Dolomite

Dolomite is important because it is generally a more permeable rock than limestone and forms
many of the best carbonate oil and gas reservoirs. Dolomite forms by the process of
dolomitization, by which CaCO3 polymorphs, such as aragonite or calcite, are replaced by
CaMg(CO3)2, or dolomite. According to Land (1983), most dolomite forms by replacement of a
precursor carbonate, whereby fluid imports Mg++, dissolves the precursor phase precipitates
dolomite, and exports Ca++.

Although some evidence for "primary" dolomite exists, many workers believe that modern
dolomite is penecontemporaneous, growing by replacement of aragonite or calcite in unindurated
muddy sediments in the supratidal zone of restricted, evaporitic tidal flat environments. Seawater
provides the magnesium to the supratidal area by two mechanisms: storm recharge and
evaporative drawdown (Hsu and Siegenthaler 1964). Such dolomite is universally composed of
small crystals ranging in size from 2 to 4 microns (microcrystalline). Because of its small grain
size and susceptibility to cementation by aragonite and Mg-calcite, supratidal dolomite does not
generally form good oil and gas reservoirs. However, dolomites formed in the intertidal and
subtidal zones of the tidal flat setting may make excellent petroleum reservoirs.
Dolomitization may also take place in thick sequences of normal subtidal marine limestone of the
middle shelf and shelf margin as well as in ancient lithified carbonates during deep burial. Good
secondary porosity may be produced in both cases. However, the dolomite formed in these
situations is the result of diagenetic processes. Thus, unlike dolomite produced in the tidal-
flat/shelf-interior setting, it is not, by itself, a reliable indicator of depositional environment.

Figure 1 shows the general depositional settings of the various carbonate lithologies and grain
types across a carbonate depositional model.

Analytical Methods and Procedures

Use of Well Logs

Many traditional well logs may not be as discriminating in carbonates as in sandstones. There are
several reasons for this:

• Carbonate successions tend to be monomineralic (or bimineralic, considering


dolomite facies); argillaceous beds and pore-fillings tend to be less frequent in
carbonates than in clastics.

• Tightly cemented skeletal lime grainstones may look just like tight lime
mudstones on many logs; both are composed of impervious CaCO3.

• Most logs do not detect high variability in pore distribution and size, or
unconnected pores.

• Acoustic logs may be much less diagnostic of porosity in carbonates than in


sandstones because grain-to-grain contacts are more cohesive, providing
misleadingly lower transit times.

As a general result, successful petrophysical synthesis of carbonate stratigraphy requires much


closer sampling intervals and discriminatory lithologic analysis. It is necessary to utilize cuttings,
cores, thin sections, x-ray diffraction and scanning electron microscopy, more detailed log study,
and more thorough seismic investigations, including vertical seismic profiling.

Importance of a Stratigraphic Framework

Satisfactory understanding of carbonate facies distribution often requires the construction of a


documented stratigraphic framework. The recommended procedure is summarized as follows:

• Construct stratigraphic cross sections, using combined sample/core logs and


wireline logs. Such sections should be oriented normal to perceived depositional
strike.

• Construct analogous cross sections parallel to depositional strike.

• Plot available fossil data, as well as possible timestratigraphic faunal markers.

• Array generalized rock types against the preliminary correlation framework.


• Adjust the correlation framework in light of information from any wells or
outcrops lying between the lines of cross section.

• Examine the distribution of primary depositional textures and diagenetic


features: Do they make sense?

• Continue to adjust, reiterate, and refine the working model.

• Commonly, many of the thickness patterns, facies distributions and diagenetic


fabrics that initially puzzle the geologist will literally "solve themselves" and
become self-evident.

Use of Seismic Stratigraphy

Because of their more pronounced cementation, erratic pore and vug distribution, and
monomineralic composition, carbonate rocks are generally less susceptible to stratigraphic
seismic modeling than their terrigenous clastic counterparts. However, the prominent depositional
topography inherent in carbonate shelf margins and pinnacle reefs, large-scale accretion
bedding, and dramatic lithologic contrasts between carbonates and salt or shale basin-fill can, in
some cases, express themselves diagnostically and dramatically on seismic sections. As in so
much of carbonate exploration, it depends on the specific geologic setting, and much empirical,
repetitive effort is required to evaluate the effectiveness of seismic exploration in any given
carbonate play.

Important Carbonate Depositional Models


The most common carbonate depositional model is the carbonate shelf. This model consists of a
wide, flat, very shallow depositional plain with evaporites and muddy dolomitic sediments formed
in the shelf interior and light-colored, coarsergrained, skeletal, calcium carbonate sediments
deposited on the shelf margin. This plain terminates seaward at an abrupt break in slope where
the depositional surface declines into deeper water of a starved basin in which sediment
accumulation has been much slower and subsidence has been greater. Basin sediments tend to
be dark and finer grained. Figure 1 and Figure 2 are a map and a profile showing the
distribution of major depositional settings of the carbonate shelf model.

A second depositional model, the carbonate ramp, was proposed by Ahr (1973). As displayed in
Figure 3 (Carbonate ramp profile), no prominent break in slope is present in this model.
Facies belts tend to broaden and "higher energy" zones tend to be relatively close to shore.
Modern and ancient examples of carbonate ramps are much less common than the carbonate
shelf model and, in fact, probably represent the earliest depositional stage in development of a
typical carbonate shelf. Depositional environments presented in the following text are, therefore,
described in the context of their setting on the carbonate shelf.
Coastal Setting

Beach

Figure 1 or Figure 2 (Generalized beach model illustrating the most simple and common
beach setting) shows the subenvironments of the generalized regressive beach model,
together with characteristic electric-log patterns, grain size, sorting, lithology, sedimentary
structures, and depositional processes. Aside from the tendency for development of cemented
dune sands, there are no intrinsic differences in process or geometry between carbonate beach
sands and terrigenous beach sands. Sand grains can be of many types, depending on the nature
of the beach and on available materials — that is, the kind of carbonate-secreting organisms that
live in adjacent shallow marine waters.
Foreshore (beach) sands merge shoreward into dune sands and seaward into shoreface sands.
Foreshore sands are characteristically coarse grained, well sorted lime grainstones, with parallel
laminations dipping gently seaward at 5º to 15º.

Shore face sands represent the subaqueous, seaward-sloping part of the beach front. Shoreface
sands are finer grained and less well-sorted than foreshore sands, and tend to range from lime
grainstones to packstones. Most common sedimentary structures are trough crossbeds
generated by longshore and tidal currents moving parallel to the beach front.

Offshore sediments represent the seaward "toe" of the beach, toward which the beach is
accreting. Offshore sediments, fine-grained and less well-sorted than shoreface sands, occur as
packstones and wackestones. Such muddy sands commonly contain horizontal-branching
burrows and more abundant marine fossils.

Carbonate sands deposited on beach foreshores had very high initial porosity and permeability.
Unfortunately, such attractive porosity is often preferentially occluded by subsequent diagenetic
effects.

Tidal Flats and Sabkhas

Tidal flats form in shallow, protected areas in the lee of physical barriers, or along the margins of
low-lying coastal areas. Sabkhas are the arid counterparts of carbonate tidal flats that form in
more temperate climates. Block diagrams from Shinn (1983) illustrate the scale and geography of
two modern tidal-flat complexes ( Figure 3 , Block diagram schematically showing major facies
on the Andros Island onlap (transgressive) tidal flat model)and Figure 4 , Block diagram
schematically showing major facies on Persian Gulf Trucial Coast offlap (regressive) tidal flat
model).
Subtidal: that part which is always immersed. The sediments on the seaward edge consist of
shelf interior, burrowed lime muds, and pellet wackestones. In the interior of the tidal flat, subtidal
sediments consist mostly of channel deposits, commonly pellet muds and mollusk-shell
packstones, that formed as placer or "lag" deposits.

Intertidal: that part of the tidal flat exposed at low tide but immersed at high tide. It consists mostly
of laminated mud in "ponds" occupying broad areas between channels and pelleted mud
deposited on the gently sloping sides of channels.

Supratidal: The largest part of the tidal flat is exposed except during storm tides. In temperate
areas, most supratidal area consists of marshland, whose arid counterpart is a sabkha.

Diagnostic features include:

• mud cracks — mostly supratidal; not preserved in intertidal zone.

• laminations — from storm layers; produce crusts; mostly supratidal.

• algal structures — stromatolitic features.

• fenestral texture — irregular pores produced by dessication.

• flat-pebble conglomerates — thin layers made of flat clasts or crusts.

• dolomitization — especially fine-crystalline dolomite muds tone.

• nodular or irregularly layered gypsum or anhydrite — associated with


dolomitized mudstone.

• vertical sequences ( Figure 5 , Typical Edwards tidal flat cycle) may differ
depending on transgressive or regressive nature of sedimentary cycle; look for
associations of above features.
Excellent petroleum reservoirs can be formed in dolomitized intertidal, subtidal, and tidal-flat
successions which provide widespread zones of fine-intercrystalline porosity. Evaporite layers
serve as top seals, seat seals, or lateral seals. An example is the San Andres formation in West
Texas where intertidal and subtidal dolomitic mudstone reservoir facies are sealed by tight
anhydritic supratidal dolomites. ( Figure 6 , North-south cross section showing three rewervoirs
sealed by impermeable anhydritic supratidal facies to the north. )
Shelf Setting

The shelf setting constitutes a broad area of shallow-marine sedimentation characterized by a


diversity of depositional environments, and most sediments show no individually definitive
characteristics indicating their origin ( Figure 1 , Facies map of the Andros lobe of Great Bahama
Bank based on optimum classification procedure. Heavy dots show location of 200 sample
stations).

Accordingly, the satisfactory determination of depositional environment ordinarily depends


upon an association of faunal, mineralogical, sedimentological, and sequential criteria, termed a
"facies mosaic" by Laporte (1967).

Of all depositional settings, shelf settings cover the largest area of modern carbonate deposition
and contain the greatest volume of ancient carbonate sediments composing the geologic record.

Muddy shelf sands are the most common depositional fabric of the shelf, typically widespread
and monotonous deposits of packstones and wackestones, possibly dolomitized. Figure 1 (A map
of modern sediment distribution on the Great Bahama Bank) shows pellet and skeletal mud
accumulations in a broad area to the west, in lee of Andros Island.
Tidal deltas typically are deposited as lime muds on the "calm" side of narrow passageways
between barrier islands. Figure 2 (Tidal deltas of upper Florida Keys) shows the location of
modern tidal deltas in southern Florida, and Figure 3 (Geological profile and environmental
subdivisions of lower Florida Keys) shows the profile across one of them.

Mud banks develop as wave-resistant features standing above the sea floor fringed by belts of
coral and algal communities on the windward side. In the rock record they might be called
bioherms or reefs, and may contain porous reservoir bodies. Figure 4 (Aerial photograph of
Rodriguez Bank ) and Figure 5 (North-south cross section, Rodriguez Bank) show a classic
example of a modern carbonate mud bank in South Florida.
Patch reefs, in modern carbonate environments, are localized concentrations of biotic
communities, dominated by coralline algae and coral and characterized by abundant and rapid
construction of skeletal/colonial carbonate debris. Typically they may range in diameter up to
perhaps 1000 in. In the geologic record, they may be short lived, leaving only lenses or patches
of skeletal sediment, or they may continue for substantial periods, creating sea-floor prominences
and/or facies-mounds that can form small but prolific petroleum reservoirs.

Figure 6 (Distribution of patch reefs and linear reefs on Great Bahama Bank) is a map showing
distribution of patch reefs on Great Bahama Bank, and Figure 7 (Simple Middle Silurian
mound near Georgetown, Indiana, in clastic-free belt of Lowenstam. Original relief low,
composed merely of micrite with bryozoans and crinoids. Lower diagram not to scale. Mound is
about 2 km across and 40 m thick. Core holes outline mound through glacial drift and partly under
thin cover of Devonian Jeffersonville and Kenneth-Kokomo formations) is an ancient example of a
patch reef in the Silurian of Indiana.
Shelf sand bodies are carbonate sand bodies, frequently extending shelfward for 10 to 15 kin
from the shelf margin, thus reaching into the general shelf setting. In most cases, such sand
bodies can be related to their shelf-margin origin.

Shelf basins are tectonic depressions, or "sags," that develop within a large shelf area. The shelf
basin itself may be filled, at least partly, with marine rocks, reflecting a deeper-water origin, or
even with euxinic shales or carbonates that can be petroleum source rocks. Reservoir quality in
shelf-basin successions is characteristically poor. They are more important as source-rock
locales and because porous shelf facies rocks may occur at their margins.

The Lower Cretaceous Maverick basin of southern Texas, a characteristically large carbonate
shelf basin (15,000 sq. mi) is illustrated in Figure 8 (Paleogeography of Middle Cretaceous
(Albian) Edwards formation of Texas. Major structural and regional depositional features and
distribution of lithological types are indicated (Cook, 1979). Broad middle shelf environments are
bounded on the southeast by the shelf-edge Stuart City reef trend.)and Figure 9 (Generalized
stratigraphic cross section showing geometric and lithologic relationships along line of section
ABC).
Shelf Margin Setting

Figure 1 (Carbonate shelf profile), a schematic profile across the carbonate shelf model,
highlights the facies developed at the shelf margin setting. These include reefs and sand bodies,
both of which can form important petroleum reservoirs.

Ecologic Reefs
This type of "reef," the result of rigid, wave-resistant organic buildups, flourishes on the shelf
margin. Modern reefs in this setting show a characteristic lateral zonation relating to depth and
wave energy levels
( Figure 2 , Cross section through a zoned marginal reef illustrating the different reef zones and
environment of different reef-building organisms). Figure 3 (A sketch illustrating the growth
form of reef-building metazoans and the type of environments in which they most commonly
occur) show characteristic shapes and morphologies of their ancient counterparts and their
respective environmental attributes. Such paleontologic data can become extremely important
to the carbonate explorer.

Shelf-Margin Sand Bodies


Ecologic reefs do not develop along all sectors of a shelf margin. Similarly, there may be times
during the existence of shelf margins, not particularly conducive to diversified organic production,
when carbonate sand shoals typically develop along shelf margins. Figure 4 (An idealized
isometric diagram illustrating facies on a carbonate shelf or platform at times in geologic history
when the only skeletal invertebrates with reef building potential were delicate ramose or
encrusting forms, and hence reef mounds are the only types of buildups. The platform margin
is a series of carbonate sand shoals and the reef mounds are restricted to the tranquil inner parts
of the platform and downslope in deeper water in front of the platform )and Figure 5 (An idealized
isometric diagram illustrating the facies on a carbonate shelf or platform at times in geologic
history when a full spectrum of reef-building skeletal metazoans is present. The platform
margin is a zoned barrier reef and isolated patch reefs dot the shallow platform behind. Reef
mounds are found on the inner part of the platform and on the fore-reef slope ) illustrate these two
end-points; commonly, however, shelf margins are marked by both ecologic reefs, as well as
shelf-margin sand bodies.

In general, shelf-margin carbonate sand bodies have two different modes of origin:

• skeletal debris — mostly washed shelfward by prevailing currents and storm


surges, but also filling passes in the shelf margin and fringing reefs.

• oolitic sands — either as elongate sand bars or belts of sand bars, just
shelfward of the shelf margin in the high-energy/ high carbonate- precipitation
zone.

Stratigraphic Detection of Shelf Margins


The characteristic subsurface pattern of a carbonate shelf margin is the presence of a significant
thickness of shallow-water carbonate rocks, which appear to change laterally over a relatively
short distance in the basinward direction to darker, often partly terrigenous carbonate rocks
having a probable deeper-water origin ( Figure 6 , Characteristic lateral relationships of shelf
margins in subsurface). The apparent (or geometric) lateral equivalence is, in most cases,
misleading; the basinal-rock unit B, although appearing to be stratigraphically equivalent to the
shelf rock A, usually turns out to be a slightly younger basin-infill unit, reflecting basin starvation
and reciprocal basin sedimentation. In other words, almost all of rock unit A is older than almost
all of rock unit B.

Seismic Detection of Shelf Margins


Carbonate shelf margins are often quite apparent on seismic sections. Figure 7 (Seismic
reflections. Upper cretaceous carbonate shelf margin, Aquitaine basin, offshore France ) and
Figure 8 (geological interpretation)show seismic data and geologic interpretation, respectively, of
an Upper Cretaceous carbonate shelf margin in the Aquitaine basin, offshore France. Note
particularly the geometric relationship between basinward-thinning, shelf accretion beds and the
basinal beds that abut the shelf margin. Figure 9 (Seismic criteria for recognizing carbonate
buildups. Criteria for directly outlining buildups include reflections from top and sides of
buildups and onlap of overlying reflections onto buildups (1-A); and patterns of seismic facies
change between buildup and enclosing strata (1-B). Criteria that indirectly outline or infer
presence of buildup include drape, velocity anomalies, and spurious events (11-A), and
determination of optimum basin positions for buildups (11-B). ) shows a variety of patterns used
for distinguishing carbonate buildups on seismic sections (Bubb and Hatlelid 1977).
Foreslope Setting

Two general types of foreslope models are recognized (McIlreath and James 1978). The
depositional type (Figure1, Schematic model for a reef-dominated depositional carbonate
margin/foreslope with sequence of deposits from slope accretion) is characterized by slopes of
sedimentation that, in turn, build gentler, accretionary slopes that merge gradually with basin
floors. The bypass type (Figure2, Schematic model for a reef-dominated bypass type of
carbonate margin, with sequence of deposits from slope accretion) is generally shallow to deep
water without much deposition on portions of the slope. Depositional models in the foreslope
setting include:
• Talus and Breccia Fans-- Coarse-grained carbonate sedimentary debris extends in
submarine fans downslope from the steeper upper parts of the carbonate foreslope.
Typically, the clasts may range upward from sand-size skeletal fragments to "house-size"
blocks of submarine-cemented reef rock contained in a mud matrix. They may extend
basinward several kilometers from the shelf-margin front. Bedding is commonly
disorganized.

• Pinnacle Reefs-- Pinnacle reefs are patch reefs, generally seaward of the main shelf-
edge barrier reef, whose column-like shape results from their location where higher
subsidence rate allows reef growth to continue. Figure 3 (Generalized cross section of
Niagara and lower Salina carbonates and evaporites, northern Michigan basin,
showing the three basic vertical lithologic successions) illustrates the location of pinnacle
reefs basinward of the shelf margin on the north side of the Michigan basin.

• Turbidites-- Water-supported, downslope gravity sedimentation can extend for


kilometers basinward from the shelf margin. Like their terrigenous counterparts,
carbonate turbidites comprise fining-upward detrital sequences associated with debris
floor sediments and dark muds with pelagic forams. Carbonate turbidites generally do not
make good reservoir rocks because of high mud content and absence of dolomitization.

Basinal Setting

The basinal setting extends from the "toe" of the carbonate foreslope seaward across a gently
declining plain to the depth below which calcium carbonate begins to dissolve in marine
sediments, generally about 4000 to 5000 in.

There are three main lithologies present in the basinal setting, as shown by studies of modern
deep-water sediments:

euxinic basinal muds — dark, organic-rich, fine-laminated, argillaceous muds; usually


petroleum source rocks.

• pelagic oozes — fine-laminated, skeletal-rich, very finegrained muds that convert to


chalk.

• turbidite sands and debris flows — coarse- to fine-grained, thin-bedded sands and
muds, with frequent slump structures.

Tectonically fractured chalks produce oil and gas in the North Sea (Lower Tertiary, Ekofisk field)
and in central Texas (Upper Cretaceous Austin Chalk).

Carbonate Case Study: San Andres Formation, Reeves Field, West


Texas

A detailed subsurface-petrographic study of the San Andres Formation, the main reservoir of the
Reeves field, West Texas ( Figure 1, Regional location map showing Permian Basin province
boundaries, San Andres production and Reeves field), was conducted by Chuber and Pusey
(1985). The purpose of the study was to gain a better understanding of the formation's
depositional history, thereby helping in the planning of infill drilling and secondary recovery
operations.

Method of Study

Chuber and Pusey first constructed numerous cross sections across the field in order to
demonstrate bed equivalency and define sedimentary and productive cycles. Correlations were
made primarily using thin (1-3 ft) shale beds with high gamma ray readings that could be traced
throughout the field.

Then, nine wells in the center of the field were selected for petrographic study. These wells had
been cored for engineering data and numerous core chips were available for study. About 1000
core chips were examined and 52 thin sections were made from selected lithologies. A slabbed
core was obtained from a dry hole, the no. 1 Smith-Brownfield, located on the extreme north end
of the field. Figure 2 is a structure map of the Reeves field, contoured on a unit near the top of the
San Andres, which shows the location of the cored well.

In describing the core chips, Chuber and Pusey's examination considered four levels of
characteristics:

1. Color, and sedimentary structures, leading to a Dunham rock texture name;

2. Proportion, size and type of grains (under a low-power binocular microscope);

3. Type and proportion of porosity, dolomite-rhomb size, and type and amount of anhydrite (under
a high-power binocular microscope);

4. Presence of fractures, pyrite, dead or live oil, earthy appearance and stylolites.

Although no chemical tests were made to determine carbonate type, slow reactions with
concentrated acid suggested the samples were entirely dolomite.

General Geology

During deposition of the San Andres, the Reeves field area is interpreted to have been a stable
shelf on the northwest flank of the Midland Basin. As typifies many San Andres fields,
hydrocarbon entrapment at Reeves results from stratigraphic porosity loss on or flanking a
structural or topographic feature. Although the map in Figure 2 indicates only about 40 ft of
maximum closure, a productive stratigraphic closure of at least 100 ft exists.

Based on lithologic and log character, the San Andres has been subdivided into four cycles,
numbered from one to four. Each cycle displays a similar lithologic gradation across the field: (1)
muddy on the southeast or basinal side of the field resulting from low-energy marine conditions,
(2) grainier and more winnowed near the crest of the field, resulting from low-energy marine
conditions, and (3) muddy on the west or shelf side resulting from low-energy shallow lagoon,
tidal flat or sabkha conditions.

A diverse assemblage of grain types were identified, with non-skeletal grains far more abundant
than skeletal grains. The most abundant rocks, however, were dolomitized carbonate muds,
recognized as having been formed from muds by the presence of worm-like tubes that survived
recrystallization.

Depositional Environment

Each cycle has a similar lithologic facies gradation indicating progressive shallowing of water
followed by a disconformity. The return to relatively deep water to commence the following cycle
is interpreted as having been "instantaneous." Based on lithologic successions, Chuber and
Pusey divided the cycles into two types: (1) shelf-edge cycles culminating with oolites and (2)
back-shelf cycles culminating with stromatolitic mud.

Back-Shelf Cycle

On the west flank of the field (shelfward), each cycle begins with organic shale overlaid by
dolomitized mudstone that grades upward into dolomitized wackestone containing fossils and
pellets. The wackestone, in turn, grades upward into weathered stromatolites containing vertical
shrinkage cracks and birdseye structures suggestive of supratidal deposition. Each cycle then
ends abruptly with a gray, one-inch-thick weathered zone. This back-shelf cycle is thus
interpreted to be a record of an initial transgression followed by a relative shallowing of water until
supratidal conditions were reached.

Shelf-Edge Cycle
An idealized illustration of the cycles encountered near the shelf edge on the east side of Reeves
field is shown in Figure 3 (Idealized shelf-edge depositional cycle in the San Andres Formation,
Reeves field ). From a sharp lower contact, organic shale grades upward to dark gray
mudstone that gradually becomes lighter in color and increases in both grain and fossil content to
become a gray fossiliferous wackestone. Above the wackestone is a fossiliferous packstone that,
in turn, grades upward into a gray or brown oolitic and fossiliferous grain-stone that caps the
sequence and contains a sharp erosional upper contact.

Based on the upward lightening of color and upward increase in fossil fragments and grains, this
cycle is interpreted to represent shoaling of water and increased water current energy as
sediment accreted in the area. Chuber and Pusey point out that, though the cycle can be easily
misinterpreted as a gradual seaward progradation, it can be distinguished as cyclic stillstand by
its multiple succession of similar cycles.

One would normally expect the oolite to be effective reservoir rock. However, postdepositional
processes have reversed the expected porosity pattern. The primary depositional porosity of the
oolite (about 25%) has been totally infilled by anhydrite. The dolomitized mudstone, on the other
hand, contains secondary leached porosity and provides for 80% of the well completion
perforations.
RECOMMENDED READING

General
Berg, 0. R., and D. G. Woolverton. 1985. Seismic stratigraphy II, an integrated approach. AAPG Mem. 39.

Bignot, G. 1985. Elements of micropalaentology. Boston: IHRDC.

Friedman, G. N., and J. E. Sanders. 1978. Principles of sedimentology. New York: Wiley.

Payton, C. E. (ed.). 1977. Seismic stratigraphy - applications to hydrocarbon exploration. AAPG Mem. 26.

Schlumberger Ltd. 1981. Schlumberger dipmeter interpretation, Vol. I — Fundamentals. New York: Schlumberger Ltd.

Selley, R. C. 1985. Ancient sedimentary environments, 3d ed. Ithaca, NY: Cornell Univ. Press.

Sheriff, R. E. 1980. Seismic stratigraphy. Boston: IHRDC.

Swanson, R. G. 1981. Sample examination manual. AAPG Methods in Exploration Series.

Walker, R. G. (ed.). 1984. Facies models, 2d ed. Geosci. Can. Reprint Series 1, Geol. Assoc. Canada.

Terrigenous Facies
Collinson, J. D., and D. B. Thompson. 1982. Sedimentary structures. London: Allen & Unwin.

Galloway, W. E., and D. K. Hobday. 1983. Terrigenous clastic depositional systems. New York: Springer-Verlag.

Harms, J. C., J. B. Southand, and R. G. Walker. 1982. Structures and sequences in clastic rocks. Lecture Notes for Short Course No. 9. SEPM.

Pettijohn, F. J., P. E. Potter, and R. Siever. 1973. Sand and sandstone. New York: Springer-Verlag.

Reading, H. G. (ed.). 1986. Sedimentary environments and facies, 2d ed. Amsterdam: Elsevier.

Reineck, H. E., and I. B. Singh. 1980: Depositional sedimentary environments, 2d ed. New York: Springer-Verlag.

Scholle, P. A. and E. Spearing (eds.). 1982. Sandstone deposi-tional environments. AAPG Mem. 31.

Carbonate Facies

Asquith, G. B. 1982. Handbook of log evaluation techniques for carbonate reservoirs. AAPG Methods in Exploration Series No. 5.

Bathurst, R. G. C. 1975. Carbonate sediments and their diagenesis. Developments in sedimentology No. 12, 2d ed. New York: Elsevier.

Ham, W. E. (ed.). 1962. Classification of carbonate rocks. AAPG Mem. 1.

Roehi, P. 9., and P. W. Choquette. 1985. Carbonate petroleum reservoirs. New York: Springer-Verlag.

Scholle, P. A. 1978. Carbonate rock constituents, textures, cements, and porosities (a color illustrated guide). AAPG Mem. 27.

Scholle, P. A., D. G. Bebout, and C. H. Moore (eds.). 1983. Carbonate depositional environments. AAPG Mem. 33.

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

Zenger, D. H., J. B. Dunham, and R. L. Ethington. 1980. Concepts and models of dolomitization. SEPM Spec. Pub. 28.
II. SEQUENCE STRATIGRAPHY
1. GENERAL CONCEPTS

Sequence Stratigraphy
Let's begin by defining sequence stratigraphy as the predictable succession of strata.

Though brief, this statement is the basis of a powerful exploration tool, one we can use to
interpret strata from outcrops, cores, well logs, and seismic data. As we learn more of the details
of sequence stratigraphy, we will refine and expand this initial definition.

Sequence stratigraphy is essential to our understanding of the marine sedimentary processes


which produce hydrocarbon source rocks, reservoirs, and seals. Sequence stratigraphy gives us
a well-defined methodology for analyzing sedimentary strata and predicting subsurface
lithologies. This methodology integrates data from a variety of exploration disciplines and tools.
Today, sequence stratigraphy, together with 3-D seismic data, is having a dramatic impact on the
economics and success ratios of petroleum exploration and development drilling.

Sequence stratigraphy is often confused with seismic stratigraphy, which represents only one of
the tools used by the sequence stratigrapher. Seismic stratigraphy, however, represents the most
important step in the development of sequence stratigraphy, and without it sequence stratigraphy
could not have evolved to its present state. We use the term seismic sequence stratigraphy to
emphasize an integrated approach to the interpretation of sedimentary strata that incorporates
seismic, well log and biostratigraphic data. The terms new or modern sequence stratigraphy may
sometimes be used to differentiate between the older, classical stratigraphy and the more
recently developed depositional sequence analysis.

The concept of sequence stratigraphy was conceived more than two hundred years ago, but its
scientific foundation developed slowly, as our understanding of Earth and its geological
processes evolved. Critical to this evolution were an understanding of the relationship between
worldwide changes in sea level and sedimentary environments, improvements in the techniques
for timing and dating geological events, and the development of sophisticated geophysical remote
sensing methods.

Stratigraphy
Stratigraphy is the geological science dedicated to the study of rock strata, their succession,
lithologies, fossil contents, distributions, and positions in time. Sedimentary rocks provide
information on the geological environments and processes responsible for their deposition, since
different depositional environments produce different sedimentary rocks. Being able to determine
depositional environments from the sedimentary record is therefore of paramount importance to
the petroleum explorationist, since some strata are more likely than others to generate
hydrocarbons and to provide the seals and reservoirs necessary for hydrocarbon accumulations.

Let's review some of the scientists whose ideas and theories contributed to the evolution of
classical stratigraphy (Figure1, History of stratigraphy from 1650 to 1950, showing important
contributors and their achievements). The following list highlights these important milestones
in geology and stratigraphy:

• In 1660, Nicolaus Steno, a Danish physician and geologist, developed the


theory that younger sediments overlie older ones, establishing the principal of
superposition.

• James Hutton, a Scottish physician and geologist, is credited as the father of


modern geology. In 1785, he formulated the doctrine of uniformitarianism,
establishing the uniformity between past and present geological processes.

• The concept of sequence stratigraphy was originally developed in 1789 by


Antoine Lavoisier. While compiling a geologic map of France in 1766, Lavoisier
developed an understanding of the relationship between sedimentary processes
and their end products. In 1789, Lavoisier published his views on stratigraphy (
Figure 2 , 1789 Lavoisier stratigraphic model). This figure is remarkably
similar to the model developed nearly 200 years later at Exxon Production
Research and which forms the basis of modern seismic sequence stratigraphy.
Lavoisier eventually drifted away from the project, and is now better known as
the father of modern chemistry. Lavoisier was sent to the guillotine following the
French Revolution for having been a tax collector for personal profit — profits he
used to finance his research.

• The principle of faunal succession and stratigraphic correlation, or what is now


termed biostratigraphy, was established in 1795 by William Smith. As a canal
engineer in England, he observed sedimentary strata and fossils, noting that the
same succession of fossil faunas could be seen in different parts of the country.
By recognizing the same fossil contents, he drew correlations betweenn
sedimentary layers in distant regions.

• Principles of Geology was published in 1833 by Charles Lyell. Strongly


influenced by the work of Hutton and Smith, Lyell divided geologic time into 13
periods, based on fossil changes between one period and the next. Lyell's
classification scheme has since undergone numerous changes and refinements.

• In 1838, the Swiss geologist Armand Gressly developed the concept of facies.
Based on his study of Jurassic outcrops in Switzerland, Gressly concluded that a
facies represents the sum total of the sedimentology and paleontology of a layer
or group of layers, and that different facies represent different depositional
environments. His definition is still appropriate today.

• The theory of a Great Ice Age was developed in 1841 by Jean Louis Agassiz.
At the time, his conclusion that erratic boulders, jumbled sediments and U-
shaped valleys in the high Alps were the products of ice sheets that once
covered large parts of the continents was startling. Attributing observed fossil
extinctions to the action of ice, Agassiz missed the overwhelming importance of
evolution. Agassiz later made important contributions to the earth sciences, and
to glacial geology in particular.

• Charles Darwin published the Origin of Species by Natural Selection in 1859.


On boarding the H.M.S. Beagle in 1831 for his famous voyage, Darwin was given
a preliminary copy of Lyell's Principles of Geology by the ship's master; Captain
Fitzroy.

While we generally attribute each of these achievements to an individual, we should realize that
many researchers made important contributions, culminating in the recognized individual's ability
to synthesize past work into an elegant solution or statement.

Stratigraphy was the dominant subject for geologists during the first half of the nineteenth century
and represented the tool for understanding the complex geology observed in outcrops around the
world. Stratigraphy enabled geologists to impose order on sedimentary deposits which initially
appeared chaotic, using the concepts of superposition, uniformitarianism and faunal succession.

During the second half of the nineteenth century, studies of the deformation of Earth's crust led to
an understanding of the importance of tectonics, but did not reach a consensus as to its causes.
Preeminent among researchers at this time was Eduard Suess, a professor of paleontology at the
University of Vienna. His analysis of mountain building and the origin of the Alps led to his treatise
The Face of the Earth (1906). In this work, he developed the concept of global tectonics. He later
proposed a theory of eustasy, or worldwide synchronous changes in sea level, to explain his
observations of the stratigraphic record.

By the start of the twentieth century, we saw the beginnings of a comprehensive picture of the
nature of Earth's crust, global tectonics and the processes of sediment transport and
accumulation. The cyclic nature of sedimentary deposits also attracted considerable attention. In
1917, Barrell published Rhythms and the Measurement of Geological Time, an explanation for
the formation of stratigraphic deposits based on cycles. Figure 3 (Sedimentary record made by
harmonic oscillations in base level ) shows how cycles operating at different time scales or
frequencies can produce cumulative results detectable in the stratigraphic record. Barrell also
developed the concept of base level — the theoretical lowest level toward which erosion of
Earth's surface constantly progresses but seldom reaches — as a relative reference point, which
would become important in the development of seismic sequence stratigraphy.

At about the same time, the Serbian meteorologist Milutin Milankovitch published his calculations
on changes in solar energy input to the surface of Earth. These changes in solar insolation are
due to orbital variations in Earth's rotation around the sun. He believed variations in solar
insolation to be the cause of observed Pleistocene climatic and glacial cycles (Milankovitch,
1920). Initially, his idea was not well-received, but was eventually verified and became an
important contribution to sequence stratigraphy, and particularly the subdiscipline of
cyclostratigraphy.

The phenomenon of cyclicity was recognized through the study of coal-bearing strata. Sharp
contrasts between sedimentary units, and their repetitive stacking patterns, indicated relatively
rapid or abrupt changes between marine and non-marine sedimentary environments. Each
depositional couplet consisted of a basal bituminous coal-bearing unit of non-marine origin
overlain by marine sediments and was interpreted to represent a transgressive/regressive marine
cycle. Weller (1930) applied the term cyclothem to these sedimentary couplets; today, the term
stratigraphic sequence would be more appropriate.

The basic concepts of modern sequence stratigraphy were in place by the early decades of the
twentieth century. The next important steps included absolute dating techniques, reflection
geophysics, plate tectonics and countless field observations.

An important breakthrough for stratigraphic geology was the discovery and use of radioisotopes
in the absolute dating of geologic strata. The potential for radioactive isotopes to accurately date
rocks had been realized soon after the turn of the century. But it was during the Manhattan
Project (1942-1946) that refined mass spectrometry instrumentation was developed, permitting
the routine analysis of isotopes. The technique provided an absolute chronology for geological
events in Earth's history a quantum leap from the previous relative methods based solely on
paleontology.

Petroleum was first discovered on land using the geophysical reflection method in 1924. By 1927,
the method was routinely applied to petroleum exploration (Dobrin, 1976). Further developments
following World War II produced the continuous seismic profiling method for marine reflection
surveys. For the first time, these tools enabled geoscientists to examine the structure and
stratigraphy of sediments underlying the world's oceans and continents. As technology improved,
new geophysical tools were developed and quickly applied to deciphering Earth's history,
particularly if these tools gave companies an economic advantage.
The development of new geophysical tools led to further investigation of several long-standing
dilemmas, including plate tectonics. The idea that all of Earth's continents had split apart and
moved was formally proposed by German meteorologist Alfred Wegener in 1912. His theory was
quickly repudiated by fellow scientists, since Wegener lacked a mechanism to explain the
movement of the continents.

Following World War II, geologists went to sea and began probing the shape and structure of the
ocean floor using more sophisticated tools. A number of discoveries quickly followed, and by
1961 Harry Hess at Princeton and Robert Schmalz at Penn State independently suggested that
Earth's mantle was convecting heat, that hot rocks were being spewed out along the axis of mid-
ocean ridges, and that the ocean floor was spreading away from these ridges and plunging under
the continents. Robert Dietz coined the term sea-floor spreading for this model. In 1967, Dan
McKenzie of Cambridge University synthesized these disparate models into the theory of plate
tectonics. This theory provided an explanation for global tectonic events, and was essential to the
study of stratigraphic sequences.

Seismic Stratigraphy
The term sequence was first applied by L.L. Sloss in 1948 to describe large-scale rock or
lithostratigraphic units, bounded by major unconformities, which extended across the cratonic
interior of North America. Sloss named his sequences after native North American tribes, and
these names still appear as mega and super sequences in the global cycle chart published by
Exxon Production Research (Haq et al., 1987). Eventually these units were recognized as the
product of cyclic changes in sea level, and the term sequence stratigraphy would develop.

During the 1950s, stratigraphy expanded rapidly from its classical base into a modern phase, in
which the interpretation of seismic reflection data would become critical. The descriptive field
methods of geology no longer provided sufficient details for exploring the subsurface in search of
minerals and hydrocarbon resources. Explorationists needed a better understanding of
sedimentary processes and deposits on a scale detailed enough to find stratigraphic traps using
remote sensing tools.

While the concept of seismic sequence stratigraphy is now generally accepted, its current
application evolved through a number of discrete steps. During the 1950s, a significant amount of
geological field work was done on modern depositional environments around the world. In
France, Gignoux published his classic text La Geologie Stratigraphique (Stratigraphic Geology).
The Huttonian dictum that "the present is the key to the past" was widely applied, This approach
led to a better understanding of basic sedimentary processes and depositional environments.
Major contributions to this research include the American Petroleum Institute's API-51 project on
barrier islands of the Texas Gulf Coast, under the direction of Francis Shepard; Harold Fisk's
classic work on the Mississippi River and Delta complex; laboratory flume experiments and field
work on turbidity flows and turbidites by Philip Kuenen in Holland, whose student Arnold Bouma
would later formulate a now classical model; and numerous field studies of reefs and carbonate
platforms throughout the world by a large number of geologists who enjoyed working in the sun
and underwater.

In 1956, Cesare Emiliani at the University of Chicago studied the oxygen isotope content of
planktonic calcareous shells from deep-sea cores and discovered that isotopes could provide a
means of approximating the paleochemistry and temperature of ocean water. This
paleotemperature technique became a valuable tool in the study of paleoclimatology,
cyclostratigraphy and biostratigraphy.

Significant contributions to research were made by petroleum companies, whose laboratories


pursued a better understanding of hydrocarbon-bearing formations and how oil and gas are
generated and trapped within formations. Based on direct field observations and shallow coring
programs in both siliciclastic and carbonate provinces, these studies led to correlations between
environmental processes and depositional lithofacies and to models which illustrate Holocene
sedimentary processes.

During the 1960s, these models were assembled into depositional systems, and used to correlate
the ancient rock record in outcrops, and, more importantly, well and seismic data. Such work on
the linkage between ancient and recent deposits made it apparent that we could infer depositional
processes from ancient depositional systems, and that these inferences might enable us to
discover new hydrocarbon source rocks and reservoirs.

At this point, researchers studying cyclic sedimentation were still documenting the existence of
depositional cycles, while trying to persuade the numerous doubters that most marine sediments
were deposited in response to fluctuations in sea level. The 1960s were also the decade of
controversy over the "innovative" ideas of continental drift and sea-floor spreading. Art Meyerhoff,
then editor of the AAPG Bulletin, was a talented and vociferous anti-drifter!

In 1967, a drilling rig was towed off the coast of southern California in an attempt to drill the sea
floor. The main objective was to reach the Mohorovicik discontinuity, which marks the boundary
between Earth's crust and outer mantle. While unsuccessful in its primary mission, the operation
demonstrated the feasibility of deep-ocean drilling, and yielded a wealth of information about
marine sediments from the retrieved cores. Its potential to science evident, the drilling program
quickly found support within the geoscience community. The Deep Sea Drilling Project (DSDP)
was funded by the National Science Foundation, through a corporation of academic institutes
from the United States and overseas. One of the remarkable feats of the DSDP program, and its
1984 successor the Ocean Drilling Program (ODP), was the retrieval of continuous, undisturbed
cores of ocean basin sediments. Until this time, paleontologists had been limited to sampling
discontinuous outcrops on land and short cores from the sea floor. The new drilling program
opened an unprecedented window on Earth's history through the detailed sampling of
stratigraphic sections from the sea floor.

In the late 1960s, digital, common-depth-point acquisition methods dramatically improved the
quality of seismic reflection data. Sources also changed, from acoustic sources like dynamite, to
more practical and repetitive sources, like VibroseisTM on land and air gun arrays at sea. In
addition, the digital format enabled high-speed computer processing and provided significant
enhancements in the retrieval of weak seismic reflection signals. This new technology gave
geoscientists a high-quality remote sensing tool useful for mapping details of the subsurface. In
retrospect, the technology, while providing an amazing improvement at the time, would be
considered primitive compared to current 3-D seismic data and its manipulation on computer
workstations.

Refinements in seismic processing also revealed low-velocity amplitude anomalies in the


subsurface. These bright spots were interpreted as indirect hydrocarbon indicators of subsurface
gas reservoirs. This interpretation approach provided an alternative to petroleum exploration
based solely on structural plays.

Later in the decade, further enhancements in exploration methods included: greater multifold
coverage of subsurface points, improved seismic processing that led to inverse modeling, seismic
data displayed in color according to frequencies and amplitudes, and relative amplitude displays.
These high-technology tools made displaying and interpreting the finer details of the subsurface
easier while reducing the ambiguity associated with previous split-spread, single-shot seismic
data. By the early 1970s, the relative ease with which offshore seismic data could be acquired led
to volumes of data on both regional and detailed scales. This data would provide researchers
fodder from which the fundamentals of seismic sequence stratigraphy would emerge.
The basic concepts of seismic stratigraphy and a methodology for subsurface interpretation were
developed in the late 1960s and early 1970s by a team of researchers headed by Dr. Peter R.
Vail at Exxon Production Research (EPR). The bulk of their work was published in 1977 as part of
AAPG Memoir 26: Seismic Stratigraphy-applications to hydrocarbon exploration. With this
publication, EPR researchers placed in the public domain the basic interpretive tools, methods,
and nomenclature for determining the stratigraphy of sequences from seismic reflection profiles.
The following three statements summarize the basis of the new methodology as applied to
marine sedimentary deposits:

• Τhe depositional sequence is the basic unit for stratigraphic analysis, and we
can subdivide the configuration and geometry of seismic reflectors into
sequences defining the stratigraphic record. (See Original seismic stratigraphic
models. Figure 1 , a generalized stratigraphic section of a sequence, with the
vertical axis measured in depth and Figure 2 , a generalized chronostratigraphic
section of a sequence, with the vertical axis measured in time. ) An individual
sequence is the product of a cyclic change in relative sea level.
• Τhe position of an individual sequence's shoreward onlap, or pinch-out of
coastal deposits, can be used to assess relative changes in sea level. We can
use this information to locate an individual stratigraphic unit on a sea level-
versus-time curve, since relative sea level or coastal onlap curves summarize the
history of the formation of individual sequences ( Figure 3 , Regional seismic
section displaying sequences).

• Τhrough an iterative process of interpreting seismic records, and extracting


the relative changes in coastal onlap, we can construct a curve showing sea level
changes through time ( Figure 4 , Sea level chart, showing global cycles of
changes in relative sea level. Cretaceous cycles (hatchured area) have not
been released for publication). EPR researchers found that sea level curves from
different sedimentary basins around the world matched, a strong indication that a
global mechanism controls the formation of stratigraphic sequences.

These concepts were based on data from seismic lines on continental shelves and slopes around
the world, and on paleontologic and lithologic data from wells along these seismic lines. Each of
these elements represents a small piece in a vast and complex jigsaw puzzle, with the final image
being a relatively comprehensive, but somewhat tentative, history of geological events
determined from the sedimentary record. An especially significant result was the implication of
global changes in sea level as the principal causative element in the production of marine
sedimentary sequences.

As we might expect, seismic stratigraphy quickly met opposition. Seismic stratigraphy was based
on the assumption that seismic reflectors follow bedding surfaces across facies boundaries, and
that seismic reflectors represent time lines or isochrons. Major reflectors were interpreted to
represent geologically instantaneous sheets of time. Seismic stratigraphy also assumed that
unconformities are surfaces along which major seismic reflections terminate.

In addition, the proprietary data used to produce the global sea level curve was not made public.
This causes controversy to this day, and has been a major impediment to the acceptance of
sequence stratigraphy by many geoscientists. Arguments against seismic stratigraphy were
frequently based on misconceptions, or on too rigid applications of the model, which was just that,
a model, and not an explanation of every conceivable sedimentary deposit. Nevertheless, as
seismic stratigraphy was used by geoscientists to interpret their own data, a consensus gradually
developed. Soon after its release, Memoir 26 became a best seller and the AAPG's most popular
publication, with more than 17,000 copies sold by 1993.

Vail and his coworkers dramatically expanded Sloss's early stratigraphic sequence approach.
Using seismic data, Vail and his coworkers applied sequence stratigraphy on a finer scale, and
developed a method for subdividing seismic reflectors into sequences. The seismic sequences
recognized by Vail et al. occur with much higher frequency than do those of Sloss, correlate
globally, and are interpreted as the products of cyclic changes in eustatic sea level.

The key to identifying sequences is to recognize their bounding surfaces or unconformities. We


can most easily identify sequence boundaries when an angular relationship exists between
underlying truncated reflectors and overlying horizons that lap onto the boundary surface (
Figure 1 , Figure 2 and Figure 3 ). These boundaries can be very difficult to recognize when
under- and overlying reflectors are parallel and an angular unconformity is not present.

Once we have identified the sequences, we can construct a chronostratigraphic chart, with the
horizontal scale representing distance (usually down dip) and the vertical scale representing time
rather than depth. The upper panel of Figure 1 and Figure 2 shows a generalized stratigraphic
section of a sequence, with depth as the vertical axis. The lower panel shows a generalized
chronostratigraphic sequence, with the stratigraphic relationships shown in the upper panel
replotted with geologic time as the vertical axis. We use chronostratigraphic charts to depict the
geological and depositional history of an area. These charts, sometimes referred to as wheeler
diagrams (Wheeler, 1958), are well-suited to interpretations based on seismic sections, because
we can consider each major seismic horizon to be a reasonable approximation of a time surface
or isochron.

By applying the methods of seismic stratigraphy, geoscientists were able to identify seismic
facies, which, when coupled with inferences based on a predictive depositional model, could be
used to estimate the gross distribution of lithologies across a grid of seismic data. Well log data,
biostratigraphy, synthetic seismograms, and experience were all critical to the evolution from
seismic stratigraphy to sequence stratigraphy, from which we can derive a surprisingly complete
depiction of subsurface rock properties.

Seismic Sequence Stratigraphy


Seismic sequence stratigraphy evolved from seismic stratigraphy and integrates geological data
from seismic sections, well logs, geochemistry and biostratigraphy. With this evolution, a complex
nomenclature developed, including such terms as systems tracts, parasequences, transgressive
surfaces, and maximum flooding surfaces. Often, these terms had confusing and contradictory
definitions.

This nomenclature, however, provides a common terminology and has led to a rigorous
consideration of the interplay between tectonic subsidence and eustatic sea level fluctuations.
This interplay is particularly evident in the control of shelfal accommodation space, the space
available for the accumulation of sediments, which, as we shall see, strongly influences sequence
geometries. In the decade following the publication of AAPG Memoir 26, seismic stratigraphy was
applied at a frenetic pace by researchers and petroleum explorationists. Had EPR really
discovered a better mouse trap, and if so, could it be put to advantage by Exxon, its competitors
and the geological community at large?

The more geoscientists learned by applying seismic stratigraphy to their everyday interpretations,
the more they discovered. Worldwide feedback brought refinements to the original models, and
greatly expanded our understanding of stratigraphic sequences. The integration of well logs,
paleontology and geochemistry to the interpretation of seismic sections added information on
facies, lithologies, stratigraphic ages, paleo water depths and paleoclimate. By integrating these
additional tools, seismic stratigraphy quickly evolved into what is now called sequence
stratigraphy, depositional sequence analysis, or, as seismic sequence stratigraphy.

The basic premise of seismic sequence stratigraphy is two-fold:

• identify large-scale sequences, covering periods of several million years, on the basis of
seismic reflector geometries;

• correlate these sequences to a global sea level chart.

With input from additional geologic data, the original sea level curve was refined and a new
version released by EPR (Haq, et al., 1987). This chart is frequently referred to as the Haq curve
and should carry a release date, just as version numbers of software programs identify release
dates. Version 3.1a of the Haq curve was issued in January, 1987. This new release of the global
sea level chart included previous proprietary data, and furnished the locations of outcrops used in
its construction. This additional data enabled researchers and users to delve into the details, and
to further refine the chart and assess its validity.

Much of the biostratigraphic data came from the DSDP and ODP drilling programs, which
supplied worldwide paleontological, geochemical and petrologic data from deep-sea cores. These
programs refined the identification and dating of planktonic and benthic microfossils into more
accurate subdivisions. This refinement was greatly assisted by the input of absolute ages and
magnetostratigraphy, the timing of reversals in Earth's magnetic poles, which leave a detectable
remnant signal in marine sediments.
Many papers describing individual case studies and refinements to the sequence stratigraphic
method were published during the 1980s. In 1990, a special research conference and publication,
Sequence Stratigraphy as an Exploration Tool, was organized under the auspices of the Gulf
Coast division of the Society of Economic Paleontologists and Mineralogists Foundation (SEPM).
Other important publications of the period include AAPG Memoir No. 39: Seismic Stratigraphy II:
An Integrated Approach to Hydrocarbon Exploration (1985); SEPM Special Publication No. 42:
Sea-Level Changes: An lntegrated Approach (1988); and AAPG Methods in Exploration Series,
No. 7: Siliciclastic Sequence Stratigraphy in Well Logs, Cores and Outcrops (1990). From the
titles alone, we can see that integrating data from a variety of sources spurred the development
and application of seismic sequence stratigraphy in petroleum exploration.

Seismic sequence stratigraphy enhanced our geological and geophysical interpretations by


providing a greater understanding of geological processes and their sedimentary deposits. This
lead to an integrated interpretive technique, which included the following tools:

• high-quality, high-resolution seismic data, with greater multifold coverage of the


subsurface. These improvements made interpretation of reflector geometries clearer,
geometries which are critical to the interpretation of sequences and their component
systems tracts.

• biostratigraphy with more accurate time subdivisions, for identifying critical sequence
boundaries established on the global sea level versus time chart. The statistical
compilation of fossil abundances and diversities greatly enhanced the picking of
condensed sections.

• identification of precise paleo water depths from microfossils, information critical for
determining depositional environments.

• use of the geochemical constituents of microfossils, such as oxygen isotopes, as


additional indicators of paleoenvironments.

• correlation of electrical well logs as indicators of specific lithofacies; the signatures of


well log curves were found to characterize specific sequence components or systems
tracts.

• the production of synthetic seismic traces from well log data and vertical seismic
profiles (VSP); these synthetics could be used to precisely correlate well and seismic
data to verify interpretations, while enhancing the integration of all the above tools.

These tools increased the sequence stratigrapher's ability to interpret subsurface geology. With
the development of deepwater exploration and deep-basin plays during the 1980s, the ability of
seismic sequence stratigraphy to identify potential reservoir sands deposited during sea level
lowstands made a significant contribution to the discovery of new petroleum reservoirs. Today,
sequence stratigraphers use the above tools to produce an integrated interpretation useful on
both the regional and the prospect level.

In the 1990s, further progress in research continues to spawn conferences and publications on
the subject. Preeminent among these was a conference on European basins held in France in
1992 to launch a project dedicated to re-evaluating those outcrops which had been the basis for
the type sections used to establish the original stratigraphic classifications. Reassessing these
original outcrops in terms of modern sequence stratigraphy is worthwhile, both to verify and to
further improve the sequence stratigraphic concept.
Important recent publications include Seismic Facies (edited by Weimer and Link, 1991), and
Cycles and Events in Stratigraphy (edited by Einsele et al., 1991). In 1994, the AAPG has no less
than six publications covering various aspects of sequence stratigraphy in the editorial
process.Before presenting a description of sedimentary processes and models, let's examine
some of the driving mechanisms responsible for the formation of cyclic depositional sequences.
We'll begin with a discussion of cyclostratigraphy.
2. CYCLOSTRATIGRAPHY

Cyclostratigraphy
While there are many ways to present the concept of sequence stratigraphy, let's begin by
discussing the cyclic nature of sedimentary strata, since cycles represent a direct link between
geological concepts, sedimentary processes and the formation of stratigraphic sequences.

A conspicuous feature of sedimentary deposits is their distinctive, rhythmic stratification, which


we can attribute to periodic or non-periodic processes, or a combination of these processes. In
general, these cyclic processes are responsible for the regularly alternating beds and the larger,
repetitive units we often see in marine sedimentary strata. Cyclostratigraphy, a subdivision of
stratigraphic geology, is the study of this cyclicity, its periodicities, magnitudes, and causes. We
can define cyclostratigraphy as the study of cyclic depositional patterns in the geologic record
produced by climatic and tectonic processes (Perlmutter and Matthews, 1989).

As we learned in the previous chapter, the most illuminating concepts put forth by modern
sequence stratigraphy are the recognition of the sequence as the basic unit of deposition, and the
global nature of cyclic changes in sea level. Furthermore, changes in relative sea level are the
primary driving mechanism for the formation of marine sedimentary sequences. In essence,
Mother Nature operates an enormous sedimentary pump driven by cyclical rises and falls in sea
level. Each cycle produces a well-defined marine sedimentary package or stratigraphic sequence.

Eustasy is the global sea-level regime and its fluctuations, caused by climatic and tectonic
processes. An understanding of the cyclic nature of changes in eustatic sea level gives us insight
into the causes of stratigraphic sequences and their timing, and enables us to date and predict
the occurrences of potential hydrocarbon source rocks and reservoirs in sedimentary strata.

Types of Cycles
We classify repetitive sedimentary cycles as either periodic or non-periodic, according to the
regularity with which they occur.

Periodic cycles are generally associated with time spans of less than a million years. These
cycles produce sedimentary deposits which include lacustrine varves and the strata resulting from
orbitally-controlled variations in climate and sea level. Varves are annual variations in the
coloration and composition of sediment layers caused by seasonal changes in sediment input to
lakes. Darker layers are due to higher organic input during spring and summer, and lighter layers
are due to the decreased organic content associated with fall and winter. Because the causative
mechanisms for such deposits are usually external to the sedimentary system under
consideration, we refer to them as allocyclic mechanisms. Allocyclic processes are usually slow
and gradual, and include tectonic uplift and subsidence, and global eustatic fluctuations in sea
level.

Non-periodic depositional processes occur irregularly or are random, and may last for periods
greater than a million years. Non-periodic cycles include both very short-term and very long-term
events. Short-term events include turbidites, storm deposits or tempestites, the migration of
channels and bars, and deltaic lobe switching, while very long term events include global tectonic
processes. Non-periodic processes involve a redistribution of sediments within a confined
system, so we refer to them as autocyclic mechanisms, which require no change in energy or
sediment input.
It is often difficult to distinguish periodic from non-periodic mechanisms, since what may first
appear to be a non-periodic pattern of discrete events may actually be the complex summation of
several cyclical factors occurring regularly over a long period of time. Cycles associated with the
formation of stratigraphic sequences have periods that span wide ranges of time, and shorter
cycles may occur within larger ones.

In summary, Nature generates periodic and non-periodic processes, both of which provide a
basis for the study of repetitive mechanisms which produce stratigraphic sequences. Now, let's
look at the details of cyclic mechanisms.

Duration of Stratigraphic Cycles


In Memoir 26 Vail and his coworkers developed a global cyclicity chart, which was subsequently
updated and revised by Haq et al. in 1987. The Cenozoic portion of this chart encompasses the
past 65 million years and shows cyclic changes in relative coastal onlap and in relative eustatic
sea level.

The cycles of relative coastal onlap in this chart represent transgressions and regressions of sea
level across continental shelves caused by relative changes in sea level. Each such cycle
produces a stratigraphic sequence. The cycles and associated changes in relative sea level were
determined from seismic stratigraphic analysis and biostratigraphic dating of marine seismic
reflections. In Figure 1 (Hierarchical scale of coastal onlap patterns (orders)) we see that larger
supercycles are composed of shorter cycles, which are composed of even shorter paracycles.
We see a hierarchy of cycles within cycles, just as classic Russian dolls are nested one into
another, or fractal patterns diverge into ever smaller entities.

The cycles were arbitrarily classified into orders based upon their duration and magnitude of
relative change in sea level (Figure1). Vail and his coworkers originally recognized three major
orders: cycles of the first, second and third order have durations of 200-300 million, 10 to 80
million and 1 to 10 million years, respectively. By the late 1980s, additional studies added a
number of shorter duration cycles (see Table1 below).
ORDER DURATION TYPE CAUSES
1st 100s ma Wilson - Megacycle 1st
2nd 10s ma Sloss - Supercycle 2nd
3rd 1-10 ma Vail - "fundamental" 3rd
4th 100-400 ka Milankovitch 4th
5th 41 ka Milankovitch 5th
6th 19 &23 Milankovitch 6th

Table 1: Cyclic orders of stratigraphic sequences

This increase in the number of orders is due to the fact that individual cycles occur on varying
time scales, and their detection is based to a large extent upon the resolution of our observations.
In the 1970s, the most commonly observed orders of stratigraphic cycles were recognized on
exploration seismic data, which may reveal only first- through third-order cyclostratigraphic
features associated with periods greater than a million years. Stratigraphic features of this
magnitude are important in the analysis of regional sedimentary basins, particularly where we
have limited well information. Higher-order cycles were recognized later in outcrops, well logs,
cores, and on high-resolution seismic profiles. These short-duration cyclic sequences occur on
the scale of individual prospects and reservoirs, and are important to production geology (1996
Geologic Society Spec. Publication 104).

To put the magnitude of these orders in perspective, we can observe their broad range of time
spans on a logarithmic scale. Figure 2 (Cyclostratigraphic time chart ) shows how cycles can
range from small, diurnal tidal fluctuations lasting just hours to first-order megacycles lasting more
than 50 million years. The chart depicts 12 orders of magnitude, with sedimentary processes
and deposition occurring throughout the entire range.

Nummendal (1993) suggested that sequence orders be named after the researchers who
contributed to their discovery or classification: first- through third-order sequences would be
known as Wilson, Sloss and Vail sequences, while higher-order sequences associated with
orbital parameters would be known as Milankovitch sequences.

Paracycles and Parasequences

Initially, third-order stratigraphic sequences were subdivided into parasequences or paracycles


(Figure1). We define a parasequence as a relatively conformable succession of genetically
related beds or bedsets bounded by marine-flooding surfaces or their correlative surfaces (Van
Wagoner et al., 1990).

As investigations into the character of stratigraphic sequences progressed, however, the size of
the paracycle shrank, since it was discovered that parasequences actually represented higher-
order sequences (e.g., fourth- or fifth-order). One important distinction between a stratigraphic
sequence and what was originally defined as a parasequence was that the parasequence did not
display an erosional unconformity related to a detectable fall in relative sea level.

While the original meaning of parasequence is still widely used, the term may more accurately
represent smaller cyclic stratigraphic units, and is best defined as the shortest detectable
subdivision within a stratigraphic sequence. We will discuss the subject of parasequences later
within the context of facies analysis.

Events
An event is any incident of probable tectonic significance, suggested by geologic or other
evidence, whose full implications are unknown. Many sedimentary processes fall into the realm of
events, including slumps, turbidity flows, floods, tsunamis, volcanic ashfalls, extraterrestrial
impacts, and storms. Although these processes are essentially instantaneous, they may
contribute significantly to the marine stratigraphic record (Figure 1, Cyclostratigraphic time chart).

We may even interpret their cumulative effects to be the result of a cyclic process or mechanism.
For example, if a coastline experiences a major hurricane every ten years, each producing an
individual tempestite, the sedimentary record would display a repetitive pattern hinting at some
underlying cyclic mechanism, which in effect it is, although not a regularly occurring one. The
same interpretation holds true for deep-sea turbidites. The seismic sequence stratigraphic
method, however, places such events in their proper perspectives and positions within an overall
stratigraphic framework. Fortunately, we can narrow our study to stratigraphic cycles of the first
through sixth orders (see Table 1 below).

ORDER DURATION TYPE CAUSES


1st 100s ma Wilson Megacycle 1st
2nd 10s ma Sloss Supercycle 2nd
3rd 1-10 ma Vail "fundamental" 3rd
4th 100-400 ka Milankovitch 4th
5th 41 ka Milankovitch 5th
6th 19 &23 Milankovitch 6th

Table 1:Cyclic orders of stratigraphic sequences


We must, however, bear in mind the large number of cyclic natural processes which leave their
track in the stratigraphic record, and that the resolution of the tools we use to detect these cycles
will limit our interpretations.
3. THE MAKING OF A STRATIGRAPHIC SEQUENCE

Inierglacial Warm Period

Sea Level Highstand (HST)

The stage is set 125,000 years before the present during the Sangoman interglacial period. The
location is the Texas-Louisiana coastal plain, shoreline, west Florida carbonate platform, and the
northern Gulf of Mexico Basin. Figure 1 (Physiography of the northern Gulf of Mexico) illustrates
the physiography of the study area. The physical and climatic environments are essentially
identical to those of today. Evidence for a higher Sangoman sea level stand is well-established
along coastlines on stable continental margins. Along the Florida peninsula, erosional bluffs
indicate that the Sangoman paleoshoreline was several meters above present sea level (
Figure 2 , Sea level curve for highstand systems tract (HST) phase).
The sequence depositional cycle begins near the end of the Sangoman interglacial period.
while this starting point is arbitrary, it is a very effective marker when associated with erosion, and
represents the sequence boundary of the Vailian model used in this example. We could, however
just as easily have selected the Sangoman condensed section to represent the cycle boundary,
and used the genetic depositional model proposed by Galloway (Xue and Galloway, 1993). The
choice of a boundary for the beginning and ending points of a stratigraphic sequence is a matter
of individual preference, since the units repeat themselves over and over. Erosional bedding
planes are usually very evident on seismic reflection profiles and to some extent on well logs,
while condensed sections are more evident on biostratigraphic abundance curves. The type of
data we use, therefore, will constrain our choice of a starting point.

During the Sangoman interglacial period, sea level is slightly higher than today, while climatic and
environmental conditions are equivalent. Coastal areas at the time are exposed to weathering,
and produce soils and fluvial drainage systems recognizable in the stratigraphic record.
Sediments accumulating during this period of the cycle form the Sangoman highstand systems
tract (HST) ( Figure 3 , Model of highstand systems tract).

During the HST stage, there is very little deposition on the outer shelf or beneath the deeper
waters of the continental slope and abyssal plain. Active sediment accumulation is restricted to
estuaries, small coastal deltas, and the Mississippi River Deltaic complex. Except for the
Mississippi deltaic complex, these areas represent small-scale sedimentary depocenters. The
Florida carbonate platform experiences the sporadic growth of patch reefs along the shoreline,
the Keys, and offshore areas, including the Florida Middle Grounds Bank. Elsewhere, deposition
consists of the slow rain of microscopic shells of plankton, and wind-transported, fine-grained
sediments, such as pollens. Depending on the prevailing winds, episodic volcanic eruptions in the
Mexican Sierras occasionally blanket a portion of the study area with a few millimeters of ash,
providing a convenient time marker in the evolving stratigraphic record.

The slow accumulation of offshore sediments during this phase of the cycle is condensed relative
to time. The Sangoman condensed section reflector drapes most of the study area, and may be
considered an isochron since it represents age-equivalent sediments. We can identify condensed
sections on high-resolution seismic records as regionally continuous, medium-amplitude
reflectors. Sediment accumulation seaward of the coastline is characterized by abundant
microfossils associated with the condensed section. These fossils provide markers for identifying
depositional environments and water depths. We can use the species, if restricted to the geologic
period under observation, to identify the stratigraphic unit of deposition. The chemical
composition of the fossils also reflects their environmental habitats. For this period, the isotope
ratio in the shells indicates warm surface waters and the absence of large volumes of continental
ice.
The correlations between isotopically derived paleo sea-surface temperatures, ages, and sea
level are based upon studies of reef terraces exposed by tectonic uplift in Barbados and the
southwestern Pacific (Farrel and Prell, 1989; Chappell and Shackleton, 1986; Bard et al., 1990;
Priazzoli et al., 1991). During periods of slow sedimentation, surficial sediments in most
environments become extensively reworked by organisms through bioturbation.

Exceptions to the low depositional regime on the shelf include the growth of shelf-edge margin
carbonate reefs. These are generally situated atop salt diapirs and a few miles landward of the
shelf edge, such as the present Flower Garden Banks located more than 100 miles offshore.
Because reefs grow in place, they are referred to as authigenic, and are not truly part of the
clastic sedimentary regime. The reef flanks, however, will form small accretionary wedges of
coarsegrained carbonate detritus. These wedges can form excellent petroleum traps when
properly sealed.

Sea level reaches its maximum shoreward advance during the maximum flooding stage. The
resultant depositional surface is the maximum flooding surface (mfs) (Figure 3 ). The maximum
flooding stage occurs during the warmest periods of the Sangoman interglacial period.

During the formation of the Sangoman highstand systems tract, the large deltaic complexes
dominate sediment input. The situation, therefore, is not much different than today, with the
Mississippi River the major contributor of sediments, and located very close to its present position
(Mcfarland and LeRoy, 1988). The deltaic complex produces a succession of individual deltaic
lobes on the shelf. The resultant stratigraphy includes repeated prograding units that overlap one
another and occasionally erode parts of previous lobes. A slight rise in sea level causes the
complex to temporarily backstep or switch laterally. In contrast, a small drop in sea level offers
less accommodation space for the delta, forcing it to prograde further onto the shelf. A transect
through this depocenter resembles the HST (Figure 3 ). Elsewhere around the basin, sediment
accumulation during this period is represented by a condensed section.

Reservoir-quality sands during the Sangoman HST are limited to those environments
experiencing dynamic sedimentation where sands can be transported, sorted and deposited.
These include shorelines, river mouths and, above all, the fluvial deltas.

The HST portion of the cycle is completed when sea level drops sufficiently to place the shoreline
along the middle portions of the continental shelf. As the cycle progresses, formerly submerged
areas of the shelf become exposed, enlarging the existing coastal plain areas. These newly
exposed areas are similar to present-day, low-lying areas of south Texas, Louisiana, and Florida,
except that the prevailing climate is slightly different. Exposed portions of the former shelf
become desiccated due to exposure to the atmosphere, erode due to rainfall and streams, and
develop a suite of characteristic soils. River systems dig deeper and become more entrenched as
sea level or base level drops (Figure 2 ). Across those subaerially exposed portions of the study
area, the Sangoman-Wisconsinan sequence boundary will form. A lack of sedimentation or
removal by erosion within these areas will produce a hiatus in the stratigraphic record, while
eroded portions will produce unconformities.

Meanwhile, fluvial sediments either erode or accumulate in a prograding manner onto the shelf,
over the condensed section and maximum flooding surface formed during the Sangoman HST.
Thus, if erosion has not occurred, the sequence boundary will be located immediately above the
Sangoman condensed section. If erosion has occurred, the boundary will follow the erosional
unconformity. In addition, the prograding complex will downlap onto the condensed section, a
very useful characteristic for interpreting seismic records.
The climatic changes and processes responsible for the gradual accumulation of snow and ice on
the Wisconsinan continental ice sheet occur very slowly, and alter the hydrologic regime over the
eastern part of the North American continent. These changes include increased rainfall, providing
runoff for the transportation of greater amounts of sediment, the lowering of base level, and
erosion of older sediments. Glacial activity in the Appalachians and southeastern Rocky
Mountains also affects the study area by eroding coarse material for eventual transport to the
sea. These climatic changes are reflected in the make-up of sedimentary depositional units and
by the fossil content and wealth of different pollens, indicating changes in environments and
biohabitats within the drainage basin.

The formation of the condensed section is now complete. Towards the shoreline, the condensed
section will merge with the erosional unconformity, forming the Sangoman-Wisconsinan
sequence boundary.

Global Cooling Period

Falling Sea Level (SMST)

With further climatic cooling and the buildup of the Wisconsinan ice sheet approximately 95,000
years ago, sea level drops and the shoreline slowly migrates towards the outer continental shelf
edge ( Figure 1 , Sea level curve for shelf margin systems tract (SMST) phase). Base level drops
and subaerially exposed coastal plain regions experience erosion.

While the prevailing climate tends to cool as the Wisconsin Ice Age progresses, climate
continues to fluctuate. These fluctuations include seasonal variations and brief periods of global
warming. The warming events are represented by fifth- and sixth-order Milankovitch fluctuations
on the paleotemperature curve ( Figure 2 , Milankovitch orbital parameters curves, summation
and oxygen isotope curve for the past 800,000 years) while average sea level and temperature
drop slowly during 80 percent of the Wisconsinan sequence cycle, small reversals occur during
the overall cooling trend. when long enough, these climatic changes bring about a reversal of
processes. A brief period of global warming can cause greater seasonal melting of ice and can
produce meltwater events, which can yield large volumes of water runoff and sediment loads,
changing depositional patterns.
Their presence in the sedimentary record may be indicated by the sudden backstepping of a
prograding deltaic sequence in association with an abrupt rise in sea level, or the increased
accumulation of sediments in lakes, bays and rivers. Such events generate parasequences,
which result from small, short-duration changes in sedimentation and represent facies changes
within the overall depositional sequence in progress.

During this period of global cooling, the oxygen isotope contents of calcareous planktonic fossils
are depleted of their lighter fraction (160), as this fraction becomes locked up within continental
ice sheets and mountain glaciers.

As the shoreline regresses, sediments prograde seaward onto the narrowing shelf when and
where supplied by deltas. Those areas experiencing major deposition develop shelf margin
systems tracts (SMST) ( Figure 3 , Model of shelf margin systems tract).
The SMST deposits are characterized by downlap onto the existing Sangoman-Wisconsinan
HST and condensed section or erosional sequence boundary, and by sigmoid bedding plane
facies and seismic reflectors. This is true for the present Mississippi river deltaic complex, which
has prograded all the way out to the shelf margin. With basinward progradation, faunas migrate
with the receding shoreline, so shallow-water assemblages are now living along the outer shelf,
indicating the gradual shift in environment and habitat.

Areas which experience little or no deposition continue to accumulate Sangoman-Wisconsinan


condensed section sediments. These areas will cover a large portion of the basin and its
shrinking margin. Thus, deposition is restricted to a number of small, shifting depocenters, while
most of the area experiences little deposition. If sea level rises following the formation of an
SMST unit, and no lowstand systems tract is produced, the SMST base becomes a Type 2
sequence boundary ( Figure 3 ).

The SMST depocenters frequently consist of lobe-shaped, deltaic sedimentary complexes which
tend to change position along the shoreline. Similarly, the Mississippi River delta has switched
locations approximately five times during the past 5,000 years (Kolb and Van Lopik, 1966), and is
presently at the outer reaches of the continental shelf.

During this period, reservoir-quality sands will be concentrated at points of fluvial discharge as
they prograde across the shelf, and along the basinward-migrating shore facies. Reefs along the
outer shelf become exposed and subjected to weathering and chemical alteration.

When sea level drops sufficiently to place the shoreline and sediment sources close to the shelf
edge, sediments are dumped down the steeper inclines of the upper continental slope to form the
lowstand systems tracts (LST).

Glacial Maxima Sea Level Lowstand (LST)


Approximately 20,000 years ago, at the peak of the Wisconsinan ice age, sea level dropped to its
lowest level, approximately 150 meters below its present level (Figure 1 , Sea level curve for
lowstand systems tract (LST) phase). This drop in sea level was a gradual and uneven
process, with intervening periods of global warming causing brief rises in sea level during the
overall drop. During the glacial maxima from 50,000 to 20,000 years ago, the shoreline and
associated coastal environments were near the shelf edge, more than a hundred miles off the
majority of the present coastline.

Rapid accumulation of sediments near the shelf edge during the preceding HST and SMST
phases created the unstable conditions necessary to initiate slumping, generate submarine mass
transport processes, and incise shelf-edge canyons. The process most likely consists of a
continuous progradational outbuilding and collapse of select SMST units, where supplied by
fluvial systems. Other areas of the former shelf may experience only a minimal accumulation of
sediments and minimal erosion by weathering processes and drainage systems.

The amount of sediments dumped onto the upper slope is a function primarily of water runoff, the
quantity of sediments transported and deposited, and their stability along the shelf edge. This
stability is controlled by the physical properties of the newly accumulated sediments and their
angle of repose on the shelf. when the physical strength or cohesion of these sediments is low
and the downward pull of gravity is sufficient to overcome the angle of repose, sediments shear
and slump downslope to initiate mass flowage. Sediments are then funneled downs-lope between
topographic highs produced by salt diapirs, or within the confines of the Mississippi Canyon and
out onto the huge Mississippi Fan. These sediments frequently contain large amounts of coarse
materials, including sands and gravels, which are deposited and redeposited as a series of facies
— the basin floor, slope fan and pro grading wedge of the lowstand systems tract (Figure 2 ,
Model of lowstand systems tract).

Sediments accumulating at the shelf edge and on the continental slope during the
Wisconsinan LST phase are composed of three very distinct depositional facies, some of which
may contain considerable amounts of sand (Figure 2 ). Sediment transport down the steeper
inclines is restricted primarily to canyons and inter- domal valleys, while accumulation is greatest
within small inter-diapiric basins, immediately downslope of the shelf-edge fluvial discharge
points, and on the Mississippi Fan. As sediments are supplied, upslope basins tend to fill and
then spill downslope to fill yet another basin. Sediments carried downslope may erode the
existing condensed section, forming an unconformity and sequence boundary; or the sequence
boundary may simply be correlative with erosional unconformities on the shelf. This is a Type 1
sequence since sea level dropped below the previous shelf edge to permit erosion on the former
shelf.
High-resolution geophysical profiles through these basins clearly illustrate the resulting
depositional sequences for the past several glacial stages or sequences. For example, the high-
resolution seismic profile in Figure 3 (High-resolution seismic profile, inner-diapiric basin) cuts
through an interdiapiric basin with a salt structure on the right. An oxygen isotope-derived,
glacio-eustatic sea level curve is shown on the left. Three prominent condensed sections, each
associated with sea level highstands, mark the last three sequence boundaries of the Late
Quaternary. Each of these represents the HST portions of Milankovitch fourth-order stratigraphic
sequences for the past 360,000 years.

During the LST phase, sediment accumulation on the slope goes through a cycle, with the
sequence boundary, identified by the underlying HST condensed section or erosional surface,
first receiving turbidite sands of the basin floor fan complex (bfc) ( Figure 2 ). The bfc facies
appear as conformable, onlapping reflectors, which overlie the condensed sections on the
seismic profile, and pinch out against the topographic high produced by the salt diapir (Figure 3 ).
These basal strata indicate the arrival of massive turbidites ahead of a channel transport system
carrying sediments from the shelf edge. Each sequence shows several episodes of turbidite
deposition, which are characterized on the section by their nearly horizontal attitude and basin-fill
nature, as if in a fluid state when deposited.

Subsequent downwarping due to subsidence and the withdrawal of underlying salt eventually re-
forms the concave structure of the basin. (Note , however, that more recent studies (e.g.
Shanmugam et al. 1995) have shown that all such "fan" structures are not necessarily sandy or
formed by turbidites.

Next, the bfc complex is covered or eroded by an active channel transport system. With time, and
the proper supply of sediments, the channels build in height to form mound-like structures and
levee overbank deposits which characterize the slope fan complex (sfc) ( Figure 2 ) (Note that
this topography may also form due to slumping. It is important to look at the overall geologic
setting when determining the nature and source of these features). This process lasts either as
long as it takes to fill the basin, at which point sediments overflow into the next basin downslope
to repeat the process, or until the sediment supply is cutoff. Figure 3 illustrates the mounded
nature of the channel complex, and the layered strata characteristic of the associated levee sheet
deposits.
The thickness of the LST systems tracts may become greatly expanded where associated with
the down-faulted portion of active growth fault systems (Figure 4 , Model of expanded systems
tracts along fault plane). Both the bfc and sfc complexes can contain large volumes of sands, and
represent prime targets for hydrocarbon exploration in deepwater depositional environments.

If the supply of sediments from the shelf edge ceases or the channel system bifurcates
upslope, slow sedimentation rates dominate and produce a minor condensed section (mcs). The
high-amplitude reflector capping the sfc complex on the profile in Figure 3 represents a typical
minor condensed section. Local slumping and distal sedimentation may occasionally produce a
thickening of sediments above the sfc complex.

Over time, the fluvial sources for these sediments may meander away from the feeding point at
the shelf edge, or a rise in sea level may cut off the sediment supply to the channels on the upper
slope. At this point, the sfc complex becomes capped by a pro grading complex (pc),
characterized by distal fine grained sedimentation (Figure 2 , Figure 3 and Figure 4 ). While
composed primarily of fine-grained sediments, these units contain occasional interbedded
shingled turbidites, which are significant hydrocarbon reservoirs in the northern Gulf of Mexico
Basin. The lowstand systems tracts developed along the west Florida carbonate platform include
numerous small, incised canyons and associated lower slope fans. These features within the
carbonate province attest to the lack of runoff and sediment supply during sea level lowstands
during the Late Quaternary.

The very abrupt rise in sea level marking the end of the Wisconsinan ice age approximately
18,000 years ago completes the formation of the lowstand systems tract. Global climatic warming
causes rapid melting of continental ice sheets and mountain glaciers and produces a rapid rise in
sea level over a 10,000 year period. As sea level rises, it transgresses across the continental
shelf to form the transgressive systems tract (TST).
Global Warming Period

Rising Sea Level (TST)

With a return to warmer climates and the melting of the Wisconsinan ice sheet approximately
18,000 years ago, sea level rises rapidly ( Figure 1 , Sea level curve for transgressive systems
tract (TST) phase). The warming climate produces massive meltwater runoff carrying large
sediment loads, and causes the shoreline to transgress across the continental shelf. The process
halts the supply of sediment to the shelf edge and upper slope, where low sedimentation rates
once again dominate and begin the formation of a Holocene-age condensed section at the
present sea floor. With the rise in sea level the basin expands its margins, causing the shoreline
to shift landward and increasing the accommodation space available for sediments.
As the shoreface transgresses the shelf, it destroys and reworks previously deposited sediments
of the HST and SMST facies of earlier sequences.. We refer to this erosional surface, which
represents the continuation of theType 1 sequence boundary from the slope and basin, as a
ravinement surface. During transgression, HST- and SMST-period deltas are eroded or
overstepped and their sands dispersed as sheet sands (Bartek et al., 1990). Where supplied,
sediments form the transgressive systems tract ( Figure 2 , Model of transgressive systems tract).

The TST facies are characterized by backstepping units which onlap the weathered and
eroded portions of the shelf. This depositional base represents the Sangoman-Wisconsinan
sequence boundary, wherever prograding facies of the LST or SMST are absent. Other areas of
the shelf experiencing no fluvial discharge form a condensed section, as is the case for the
present-day shelf of the northern Gulf of Mexico. Under suitable environmental conditions, reef
growth may also accelerate on the carbonate platform areas. Reservoir-quality sands are
produced within TST units in association with the transgressing shoreline and small fluvial deltas
and estuaries.

The transgressive phase is completed when sea level reaches its highest level and maximum
coastal onlap position. Holocene interglacial climatic conditions prevail as they do today. Again,
we have a maximum flooding surface, but this time for the Holocene.

Summary
We've seen how a stratigraphic sequence represents the depositional record of an individual
cyclic fall and rise of sea level. The cyclic depositional processes described above illustrate the
development of a clastic depositional stratigraphic sequence, including the processes on the
adjacent carbonate platform. The time frame of our example makes it a fourth-order stratigraphic
sequence, containing smaller eustatic variations of the fifth and sixth order, all of which fall into
the realm of Milankovitch orbital-climatic cycles.

We should note that not all areas in this example will contain a complete succession of systems
tracts, as some areas may not receive large volumes of sediments, or erosion may remove
certain components during the course of the cycle. These factors, coupled with the regional
effects of tectonic subsidence and uplift, add to the complexity of our task in deciphering the
geologic record. We should also note that the above processes take place worldwide within all
interconnected oceans and seas affected by the same eustatic processes. Climatic changes
affect freshwater lakes to essentially the same extent.

With this understanding of a typical sequence and its components, we will now consider how we
can use seismic sections, well logs, and biostratigraphic data to recognize sedimentary
sequences. We will take a second look at what Vail and others have portrayed in the global cycle
chart, and how we can use this chronostratigraphic framework to interpret the sedimentary
record. The chronostratigraphic cycle charts and the sequence model combine to give us a
powerful predictive tool for analyzing exploration data.
4. SEQUENCE STRATIGRAPHY

Relative Sea Level Components and Accommodation Space


The marine stratigraphic record is composed of depositional sequences. The formation of these
sequences is controlled by variations in sediment supply, eustasy, compaction and subsidence (
Figure 1 , Variables that control sequence formation).

Sediment supply is a function of erosion, runoff and fluvial transport loads and processes.
Sediment supply is strongly affected by the combination of tectonics, climate and drainage basin
size. The tectonic uplifting of a land mass increases the potential for erosion and transport, even
if it reduces the size of a drainage basin. Climate also affects environmental conditions such as
aridity, precipitation, and glaciation, which in turn control weathering, erosion and sediment
transport.

Eustasy, is the global component of sea level change caused by the sum of climatic variables and
glacial and tectonic processes. Changes in relative sea level alter the base level toward which
fluvial erosional processes strive. Eustasy is affected by changes in ocean basin volumes, such
as those caused by mantle convection and swelling along mid-ocean ridges. Eustasy is also
affected by the volume of water contained within the ocean basins, which can be strongly
influenced by glaciation and, to a smaller degree, by global climate.

Compaction reduces the volume of marine sediments by expelling interstitial pore waters and is
caused either by desiccation or overlying sediment accumulation. While a relatively slow process,
compaction is quick with respect to geologic time (Figure 2 , Cyclostratigraphic time chart).
Compaction transforms loosely deposited marine clays and silts into compacted shales by
reducing volume several-fold. We should note that sands experience very little compaction
compared to clays.
Subsidence occurs on a regional scale. A typical example is the isostatic adjustment required
by the sedimentary infilling of a basin. Subsidence can be a very slow process, and is a function
of sediment supply and upper mantle processes ( Figure 2 ). The magnitude of subsidence is
highly variable but is usually on the order of meters per million years. However, tectonic
subsidence or uplift can also be a relatively rapid, high-magnitude process, similar to local faulting
and salt tectonics.
Relative sea level is the sum of eustasy, subsidence and compaction and controls the base level
to which sediments will be eroded. Relative sea level also controls the accommodation space
available for the accumulation of sediment. Sediment supply fills the accommodation space
created, thus controlling water depth. Figure 3 (Effect of subsidence on eustasy, and sequence
stratigraphic terminology) illustrates the effect of this process on eustasy and subsidence.

In Figure 3 , we see how we add eustasy, shown as a sinusoidal curve varying with time, to
the effects of tectonic subsidence, shown as a linear function, to produce a relative change in sea
level and, hence, base level. Note that while no scales are shown, subsidence and eustasy each
have rates of change which vary with time. The net effect depicted in the figure is that of a relative
rise in sea level due to subsidence, which leads to the filling of the available accommodation
space. The figure shows a general scheme, which will vary for a given cyclostratigraphic order,
the amount of subsidence, and the sediment supply. Figure 3 is appropriate for a long-term, third-
order cycle, but is not appropriate for a short-term, fourth-order sea level fluctuation, where the
effects of subsidence may be negligible. Note the use of specific line types used to delineate the
various surfaces within a sequence .

We can incorporate all of the above factors into an equation that describes the sedimentary
depositional supply system and the resulting stratigraphic record:

Accommodation Space = Eustasy + Subsidence + Compaction

In addition, the rate at which sediment fills accommodation space also affects the water depth,
whether we are considering progradational or retrogradational stratal facies:

Water Depth = Eustasy + Subsidence + Compaction - Sediment Accumulation

We can use these two equations to design computer models which depict the infilling of a
sedimentary basin. These models will generate stratigraphic sequences according to the input of
variable parameters over time. For a more detailed discussion of accommodation space, see
Posamentier et al. (1988). For examples of computer-based modeling of sequence generation,
see Rouch et al. (1993), and Lawrence (1993).

Having considered the processes which produce stratigraphic sequences, let's now look at the
resulting stratal patterns, which are so important in interpreting sequences from seismic reflection
records.

Depositional Geometries of Stratigraphic Sequences


The geometries or patterns of strata, most frequently determined from seismic reflectors, provide
the key to interpreting and analyzing systems tracts and sequence boundaries in seismic
sequence stratigraphy. The coarsest internal elements of sequence systems tracts are
parasequences. The balance between sediment supply and relative sea level controls overall
depositional geometries, which fall into the following three major categories:

• progradational

• aggradational

• retrogradational
Progradational geometries result from a dominance of sediment supply over accommodation
space (Figure 1 , Depositional geometries). This surplus causes deposition to migrate
basinward, where sediments have space. Progradation produces stratal clinoforms which display
the outward migration of the offlap break. The offlap break, or inflection point, frequently
represents the shelf edge, separating coastal plain topsets and neritic facies from upper slope
(bathyal) foresets and basinal facies.

Prograding complexes are characteristic of the shelf margin, highstand, and portions of the
lowstand systems tracts, where internal strata (clinoforms) may represent either parasequences
or low-order (i.e., fifth or sixth) cyclostratigraphic units. On well logs, an upward-coarsening profile
of sedimentary components suggests progradation. This profile results from an upward shift from
fine-grained, outer-neritic sediments to coarser-grained, shoreline-littoral depositional
environments as sediments accumulate at a given location.

Aggradational geometries occur when sediment supply matches the increase in accommodation
space caused by either rising sea level or tectonic subsidence. This balance produces vertically
stacked facies ( Figure 1 ) and an offlap inflection point that does not migrate. Electric logs
generally show repetitive units, or parasequences, of upward-coarsening patterns or mixed
components.

Retrogradational geometries result from either sediment starvation or a net increase in


accommodation space, due to a relative sea level rise, over sediment supply. Facies patterns,
including the shoreface, migrate landward. We see an upward shift from coarse-grained, inner-
neritic deposits to deeper, pelagic depositional environments, as sea level rises and the shoreline
migrates. Retrogradational geometries are associated predominantly with transgressive systems
tract deposition, and produce fining-upward electric log patterns. The transgressive systems tract
geometry of Figure 1 is the result of low sediment supply and/or a rapid rise in sea level, which
frequently results in the erosion of the continental shelf.
We can also subdivide depositional geometries according to depositional environments based on
seismic reflection character (Sangree and Widmier, 1977). We can make similar subdivisions at
the parasequence level, including beds and laminae, by integrating seismic data, well logs and
cores and using the concepts established by Van Wagoner et al. (1990). These subdivisions of
marine sedimentary rocks are outlined in Figure 2 (Stratigraphic units of marine sedimentary
rocks ), which shows the breakdown of sequence components. We begin with systems tracts,
which are subdivided into parasequences and parasequence sets, which, in turn, are subdivided
into bed sets composed of laminae and laminae sets.

Parasequences and parasequence sets define the internal clinoforms, or bedding planes, and the
stacking patterns associated with depositional geometries. We subdivide sequences and systems
tracts into parasequences using high-resolution tools, or detailed observations of outcrops. This
analysis provides the invaluable details necessary for understanding sedimentary facies at the
scale of individual hydrocarbon reservoirs.
In addition to the major depositional geometries we have just discussed, there are a number of
geometrical configurations which characterize the lowstand systems tract on seismic reflection
profiles. These include chaotic zones indicative of mass slumping, horizontal geometries caused
by the presence of turbidites within basin floor complexes, and mounded, "gull-wing" patterns
associated with channel levee systems within slope fan complexes (Figure 3 , Log responses of
different systems tract on the shelf and slope).

For those readers unfamiliar with the analysis of electric well log data and their use in identifying
lithologies within sequence stratigraphic units, please refer to standard texts on geophysical
prospecting methods or drilling industry publications. As a general rule, as water-saturated
sediments coarsen, their electrical resistivity increases and their spontaneous-potential (SP) or
gamma ray (GR) values decrease, so that the two curves produce mirror images of each other
(Figure 3 ). By integrating well logs with seismic reflection data and biostratigraphy, we can more
definitively identify the components of stratigraphic sequences.

Carbonate Sequence Stratigraphic Models


Many conceptual models illustrating the geometries and systems tract components of
stratigraphic sequences have been developed. These models show how sedimentary basins fill,
or carbonate platforms build, in response to cyclic changes in relative sea level. They are usually
idealized down-dip cross sections, which transect from shoreline to the basin center. Models
provide a valuable framework for understanding and interpreting seismic sequence stratigraphy in
both carbonate and clastic depositional environments.

Carbonate depositional systems, while prolific producers of oil and gas, are primarily organic
growth systems. Their growth and evolution, however, occur primarily during periods of rising
relative sea level. Thus, we may expect to find unconformity-bounded stages of growth linked to
eustatic control, with the unconformities forming during phases of sea level lowstand and
subaerial exposure. To date, several sequence stratigraphic models have been proposed for
carbonate deposition, four of which are illustrated in Figure 1 (Diagram of different types of
carbonate depositional profiles). Such models will eventually provide a basis for interpreting
carbonate platforms, ramps and margins.
The illustrated models are similar to those for clastic sedimentary environments, since the eroded
clastic portions of a carbonate platform will frequently duplicate the clastic models. The sigmoid,
or S-shaped, planes of the model indicate the characteristic parasequences resulting from
basinward growth and deposition. Thus sedimentation from carbonate platform rims and reef
margins provides a setting similar to that of the conceptual model presented in Figure 2 (Log
responses of different systems tract on the shelf and slope).

Figure 3 (Carbonate platform margin, southern Florida) shows a representative, high-resolution


seismic section from southern Florida. This section displays the sequences produced by eight
glacio-eustatic cycles during the past one million years, with parasequence clinoforms produced
by outward building and progradation of clastics from a stable carbonate platform. The profile also
shows well-defined erosional sequence boundaries formed by subaerial exposure and ocean
currents during Late Quaternary lowstands in sea level.
Clastic Sequence Stratigraphic Model

The most commonly used model in clastic sequence stratigraphy is an updated version of the
model originally proposed by P. R. Vail and his colleagues at Exxon Production Research (
Figure 1 , Clastic depositional sequence model). This model has evolved dramatically from the
initial model by Vail et al. (1977) (Original seismic stratigraphic models Figure 2 , shows a
generalized stratigraphic section of a sequence with the vertical axis measured in depth) and
Figure 3 , (shows a generalized chronostratigraphic section of a sequence, with the vertical axis
measured in time). The current model is frequently referred to as the "Vail" or "Vailian" model
(aka "slug" model).
The clastic sequence depositional model in Figure 1 shows how a sedimentary basin might fill
during a fall and rise in sea level. We can subdivide the sedimentary strata into a succession of
systems tracts, based on stratal geometries. For each grouping, stratal geometries characterize
sedimentation during a specific phase of a sea level cycle, as shown on the sea level curve in the
lower left of the figure. Thus, sequences contain a predictable series of linked depositional
systems, from the point of maximum shoreward transgression out to the deepwater, bathyal
environments. (Note the use of specific line types to outline the sequence boundaries and
systems tracts. The use of standards provides a common basis for communicating the results of
a seismic sequence stratigraphic analysis in reports, maps, and cross sections.)

Sequence development is driven by cyclic fluctuations in relative sea level. Systems tracts are
produced during specific phases of this cyclicity. Whatever the driving mechanism, it is clear that
cyclic changes in relative sea level, of varying orders of magnitude, control the patterns of
shoreline regressions, transgressions, and associated depositional packages we observe in the
geologic record.

The nomenclature of Figure 4 (Effect of subsidence on eustasy, and sequence stratigraphic


terminology) and Figure 1 is now widely used by geoscientists in exploration, production, and
academia. Other models and modifications have been developed for different types of margins
and depositional situations, i.e., the growth fault setting shown in Figure 5 (Model of expanded
systems tracts along fault plane). The genetic sequence model proposed by Galloway
(1989a), in which condensed sections are used to delineate sequences rather than erosional
boundaries, has now been updated (Xue and Galloway, 1993) to match the Vailian model of
Figure 1 . The choice of using either an erosional unconformity or a condensed section as a
sequence boundary is one of personal preference, and is dictated primarily by the tools used for
an interpretation and the features most evident on the available data.

Clastic model: Sequence Boundaries (SB)

Two types of sequences and sequence boundaries have been defined


(Figure 1 , Clastic depositional sequence model). The Type 1 sequence is bounded at its base
by a Type 1 sequence boundary, characterized by subaerial exposure and erosion produced by a
relative fall in sea level. We identify a Type 1 sequence boundary by a basinward shift of coastal
facies, and fluvial incision. A Type 1 sequence consists of three systems tracts — the lowstand,
transgressive and highstand systems tracts — but does not contain a shelf margin systems tract.
Type 1 sequences and sequence boundaries are the most commonly observed, and represent
the setting for those exposed portions of the continental shelf during the lower sea level phase.
Unless otherwise stated, in this text the terms sequence and sequence boundary refer to a Type
1 sequence and sequence boundary.

A Type 2 sequence has a Type 2 sequence boundary at its base, and either a Type 1 or Type 2
sequence boundary at its top ( Figure 1 ). The Type 2 sequence boundary lacks the strong
erosional features associated with the Type 1 boundary, and the basinward shift of coastal facies,
because the sea level in this cas doesn't fall below the shelf edge. This type of sequence is,
however; considered rare within clastic depositional sequences (Van Wagoner et al., 1990).

In order to extract a sequence from seismic exploration data, we must first define and trace its
sequence boundaries. We usually do this on seismic sections, tracing the discordant relations of
strata along unconformities, or; with more difficulty, along conformities. Recognizing the
systematic patterns of seismic reflector termination along surfaces, such as onlap, downlap or
truncation, is an essential step in identifying sequence boundaries on seismic profiles.
Paleontologic and well log data can be immensely helpful in selecting both sequence boundaries
and especially condensed sections, which are characterized by high high abundane and diversity
of fossils. The characteristic onlap and downlap patterns, shown in Figure 1 , provide the basic
guidelines for selecting sequence boundaries and delineating sequences and systems tracts on
seismic profiles.

The interpretation of structure, faults, erosional truncations, incised valleys and canyons from
regional seismic reflection profiles represents the first step in a seismic sequence stratigraphic
investigation. We usually do this on a regional scale, by extracting the major sequence
boundaries, and by determining paleodepositional environments from well logs and paleontologic
data. Figure 2 shows the results of such an interpretation from the northern Gulf of Mexico.

The sequence stratigraphy shown in Figure 2 is based on a detailed interpretation which


integrates electric well logs, biostratigraphy and synthetic seismic traces. Figure 2 crosses the
outer shelf and upper slope of the northern Gulf of Mexico. The structure is due primarily to the
effects of salt tectonics, with the salt depicted in blue. Ages for the major sequence boundaries
recognized in the Gulf of Mexico follow the Mio-Pleistocene sea level biostratigraphic chronology
of Figure 3 .

The simplified sea level chart in Figure 3 covers the past six million years and depicts the relative
change in coastal onlap, sequence boundary ages, and primary bio-markers used for dating
sequences in the Gulf of Mexico. This sea level chart is typical of locally produced charts and
contains refinements to the global chart of Haq et al. (1987).

Clastic model: Lowstand Systems Tract (LST)

The lowstand systems tract ( Figure 1 , Clastic depositional sequence model) is bounded at its
base by a Type 1 sequence boundary on the shelf, and its correlative counterpart above the
condensed section in deeper waters. Its top is marked by the transgressive and highstand
systems tract surfaces. The lowstand systems tract is composed of three components — the
basin floor complex (labeled as 1sf for lowstand fan), the slope fan complex (labeled as mf/co for
mounded fan and channel overbank), and the prograding complex (labeled as pgc and 1sw for
prograding complex and lowstand wedge). (Note the use of a variety of terms for describing
common features, and how this can be confusing. For the sake of clarity, all sequence
stratigraphers should adopt a common terminology.)

The lowstand systems tract is deposited in deeper bathyal environments during times of a fall or
lowstand in relative sea level. During such periods, accommodation space on shelf topsets is
reduced or non-existent due to subaerial exposure ( Figure 2 , Depositional geometries).
Lowstand systems tract deposition may continue during the subsequent rise in sea level.

Lowstand systems tract deposits display characteristic internal geometries, such as


downlapping and onlapping mounded fans, conical "gullwing" patterns, downlapping channel
levee structures, chaotic slump zones and wedges, and shingled turbidite mounds within basin fill
prograding wedges. While their overall morphologies may be strongly controlled by depositional
environments. These environments include the relatively featureless stretches of a continental
slope, or the more constrained environments associated with canyons and inter-diapiric basins,
sills, or other structural features. During the lowstand systems tract phase, incised valleys and
canyons are eroded on the shelf and at the shelf edge. These become filled during the
subsequent rise in sea level and are thus labeled incised valley fill (ivf). Downslope of these
features, the thickest accumulations of reservoir quality sands are deposited.
Sedimentary basins with well-defined shelf edges produce well-defined LST internal components,
which include the basin floor complex (bfc), slope fan complex (sfc), and prograding complex (pc)
( Figure 3 (Log responses of different systems tract on the shelf and slope ) and Figure 1 ).

The basin floor corn pier (bfc) is composed predominantly of the turbidite sands and interlayered
shales of submarine fans. These units are generally located downslope of canyon-channel feeder
systems originating at the shelf margin. The basin floor complex is characterized by seismic
reflectors with downlap and onlap terminations ( Figure 1 ). The basin floor complex typically
contains excellent reservoir quality sands which display a blocky or "boxcar" pattern on well logs (
Figure 3 ). The bfc sands frequently have high permeabilities and porosities, limited continuity,
and excellent seals when capped by pelagic shales. The bfc overlie the preceding highstand
condensed section and sequence boundary, and are capped by the overlying slope fan complex.
Minor condensed sections may occur on the tops of basin floor and slope fan complexes. These
condensed sections are indicative of a period of pelagic deposition associated with low sediment
supply ( Figure 1 ).
The slope fan complex (sfc) generally grades upwards from pelagic shales into channel transport
complexes, which may or may not contain overbank levee deposits. Sands of good reservoir
potential are found within channel axes, and as sheet sands on the levee flanks. These sands,
however; may be restricted laterally. Figure 3 shows an idealized electric log response for the
slope fan complex, with a typical coarsening-upward pattern. Repetition of this well log pattern
indicates repeated episodes of channeling. Very high electric log responses may indicate coarser
lag deposits within the channel axes. The sfc is frequently capped by a minor condensed section
( Figure 4 , High-resolution seismic profile, inter-diapiric basin). Note that faunal contents
within minor condensed sections may indicate cooler environments associated with the LST
phase, and this may serve to differentiate them from major condensed section faunas.

Deposition associated with the onset of renewed coastal onlap across the shelf produces a
prograding complex (pgc), and results in the supply of distal sediments to the LST complex on the
slope or in the deeper parts of the basin. Sand deposition is limited to shoreline areas along the
outer shelf, which will include fluvial and shoreface facies that commonly display a coarsening-
upward well log pattern ( Figure 3 ). During development of the prograding complex, turbidites
occur occasionally, producing shingled turbidite sands within thick shales. The prograding
complex is capped by deepwater distal deposits of the transgressive and highstand systems
tracts. These transgressive and highstand units are separated by the maximum flooding surface
Where deposition is limited a condensed section may develop encompassing the upper part of
the TST, the mfs and the lower part of the HST associated with the rise in relative sea level.

Lowstand systems tract deposits thus overlie highstand systems tracts and condensed section
units, and are capped by the next chronostratigraphic unit, a condensed section. The LST
subunits are colored in yellow (bfc), brown (sfc) and purple (pc) on the interpreted seismic section
of Figure 5 (Regional seismic sequence stratigraphy line: northern Gulf of Mexico Basin ).
Note the greater number of LST units in the deeper parts of the seismic section, indicating
deeper water depositional environments, and how these units grade into shallower water
depositional units shoreward (toward the right of the figure).
The bfc and sfc sands provide excellent reservoirs for hydrocarbons, while the pc unit shales are
good seals for the underlying sands. Studies of LST turbidite systems by Reading and Richards
(1994) have led them to classify these systems into twelve groups, based on grain size and type
of feeder system. Such guidelines permit a more accurate prediction of the lithologies and
reservoir potential of lowstand systems tracts.

Clastic model: Transgressive Systems Tract (TST)

With a rise in relative sea level and transgression of the shoreline across the shelf, sediment
supply to LST complexes is cut off, except for continued distal accumulation within the prograding
complex. Sediments accumulate to form the transgressive systems tract ( Figure 1 (Log
responses of different systems tract on the shelf and slope ) and Figure 2 (Clastic depositional
sequence model )). The transgressive systems tract is characterized by dipping stratal
geometries which define the retrogradational accumulation of sediments on top of the
transgressive, or ravinement, surface and lowstand systems tract at the outer shelf margin. The
internal facies, or parasequences, onlap the underlying sequence boundary in a landward
direction, and downlap onto the transgressive surface basinward. The top of the TST will contain
the maximum flooding surface and condensed section, onto which the highstand systems tract
will then downlap ( Figure 2 ).

Transgression indicates either a decrease in sediment supply or an increase in accommodation


space associated with a relative rise in sea level. Sedimentary facies at a given location within the
transgressive systems tract tend to fine upward, as a result of progressive deepening of
depositional environments and the landward shift of the shoreline ( Figure 1 ). Basinward the TST
sediments tend to thin, and we recognize them on seismic profiles by an apparent truncation
surface at their top ( Figure 1 ). Reservoir quality sands within TST units may include beach and
shoreline facies, and to a lesser extent lagoonal backshore facies. Thus, the TST unit
encompasses environments from littoral to outer shelf neritic.

During transgression, older deltaic complexes, built up and out across the shelf during the
previous SMST or LST phases, are eroded or overstepped, a process which extensively
redistributes sands as sheets across the shelf. As the rise in sea level stabilizes, the shoreline
attains its maximum shoreward incursion and a high-stand systems tract begins to form.
Transgressive systems tracts are underlain by a sequence boundary or LST unit, are capped by
an HST unit, and are colored green on the interpreted section of Figure 3 (Regional seismic
sequence stratigraphy line: northern Gulf of Mexico Basin ).
Clastic model: Highstand Systems Tract (HST)

The highstand systems tract (HST) represents the uppermost unit of a depositional sequence (
Figure 1 , Clastic depositional sequence model). It overlies the preceding TST phase and is
capped by a Type 1 or 2 sequence boundary. We recognize the highstand systems tract on
seismic profiles by downlap onto the maximum flooding surface condensed section.
The HST occurs when the sediment supply rate exceeds the accommodation space, causing
parasequence deposition to either aggrade upwards or prograde basinward. This progradation of
facies is frequently characterized by an upward coarsening of sediments, as seen on well log
patterns ( Figure 2 , Log responses of different systems tract on the shelf and slope) Reservoir
quality sands within the HST are characterized by ribbons of sheet sands produced by shoreline
facies, and by point sources resulting from fluvial deltaic depocenters. The HST units are
colored orange on the interpreted seismic section of Figure 3 (Regional seismic sequence
stratigraphy line: northern Gulf of Mexico Basin ).
Clastic model: Shelf Margin Systems Tract (SMST)

Shelf margin systems tracts are not frequently detected, since they represent a restricted period
of deposition at the outer shelf in association with a small drop and subsequent rise in sea level
or progradation of the sediment supply in search of accommodation space. The net result of
these processes is the formation of an aggradational complex, which downlaps onto a Type 2
sequence boundary at the outer shelf ( Figure 1 , Clastic depositional sequence model). The
SMST complex may be capped by either a Type 1 or 2 sequence boundary, and will not contain
an associated lowstand systems tract ( Figure 1 ).

Clastic model: Maximum Flooding Surface (mfs)

As described in association with the TST and HST systems tracts, a maximum flooding surface
develops during the maximum landward incursion of the shoreline. Condensed sections, in turn,
form within the maximum flooding surface deposits. The condensesed section associated with the
exhibits pelagic deposition and sediment starvation on the shelf and slope, and separates phases
of shoreward retrogradation (transgression) from those of basinward progradation (regression).

While the mfs is not an unconformity, and is therefore not a sequence boundary, it represents a
fundamental depositional surface within a sequence. Furthermore, it can be a prominent and
useful feature in seismic interpretation, since it provides the downlap surface for the prograding
units of the highstand systems tract on the shelf and produces a high-amplitude, regional reflector
in deepwater environments ( Figure 1 , High-resolution seismic profile, inter-diapiric basin).
The maximum flooding surface condensed section may be indicated by a small inflection on
electric well log curves. The mfs condensed section contains the primary biostratigraphic markers
we use to interpret stratigraphic sequences from well data ( Figure 2 , Regional seismic sequence
stratigraphy line: northern Gulf of Mexico Basin). The condensed section is indicated by peaks in
both faunal abundances and diversities.

Following the genetic models of Galloway (1989a), we can sometimes use the mfs, rather than
the erosion-controlled sequence boundaries, to delineate stratigraphic sequences. The maximum
flooding surface and associated condensed section sometimes provide a readily detectable
reference surface on seismic records, well logs and biostratigraphic data, which we can use to
regionally correlate stratigraphic sequences ( Figure 2 and Figure 3 (Mio-Pleistocene sequence
chronostratigraphy, northern Gulf of Mexico Basin )).

The models we have discussed so far represent conceptual frameworks which we can use to
decipher the stratigraphic record. We must, however; be flexible when making interpretations and
applying models, since every situation will be different and at variance with a particular model.
We will return to the characterization of stratigraphic sequences using well log and
biostratigraphic data.
Chronostratigraphy and Lithostratigraphy
In simple terms, chronostratigraphy is the descriptive study of sedimentary strata in terms of age.
We define a chronostratigraphic sedimentary unit by its age. In seismic sequence stratigraphy,
we can derive ages by correlating coastal onlap patterns to the global eustatic cycle chart, which
provides a chronostratigraphic framework. This process does not require the absolute dating of
strata based on geochemical or paleontologic information, and is therefore very useful when
interpreting seismic data in an unexplored basin where we may not have well data.

Clastic sediments are deposited in layers or parasequences, which consist of beds and bedsets
(Figure 1 , Stratigraphic units of marine sedimentary rocks). This stratification results from the
distribution of similar sediment types or lithologies over a given area, under similar environmental
conditions. As environmental conditions change, so do the types of sediments being deposited. In
this manner; stratal layers of differing lithologies are produced. Non-deposition generates a hiatus
in the sedimentary record. The net result of this depositional process is that, for a given location,
sediments within layers tend to be more alike than between layers.

Furthermore, surfaces separating parasequences represent an indefinite period of time — the


time it takes for the depositional environment to change. Each such surface, for all practical
purposes, represents a finite span of geological time, such that the associated sediment layers
represent time-stratigraphic or chronostratigraphic units.

Figure 2 (Depth (lithostratigraphic) profile ) and Figure 3 (Time (chronostratigraphic) profile.


) A Wheeler diagram, provides a nice visualization of depositional events through time, where
voids represent periods of non-deposition or erosion. Recent releases of global and local sea
level charts use this type of display to emphasize details obtained from type outcrop locations
around the world.
Sedimentary depositional environments, however; also vary laterally, so strata deposited during
the same time interval, but under different depositional environments, will contain different
lithologies. Strata classified according to rock types are called lithostratigraphic units, and will not
necessarily follow chronostratigraphic units.

Seismic reflections are caused by changes in acoustic impedance, usually in response to


changes in lithology. In seismic stratigraphy, we assume that reflectors follow bedding planes,
which cross lithologic boundaries, and thus represent chronostratigraphic time lines. Exceptions
occur where rapid lateral changes in lithology and lithologic facies, rather than bedding plane
reflectors, become important. Other exceptions include gas hydrate boundaries, which are a
function of temperature and depth and thus cut across strata, as do gas-oil-water contacts.

The end product of a seismic sequence stratigraphic interpretation is a complete chronology of


events and deposition, whether along a transect or across an entire basin ( Figure 4 , Regional
seismic sequence stratigraphy line: northern Gulf of Mexico Basin).
5. INTERPRETATION OF STRATIGRAPHIC SEQUENCES

Integrated Interpretation Procedures


The following is an outline of an integrated interpretation procedure, which will reveal the
sequence stratigraphy for a targeted exploration area. Using a consistent procedure results in the
organized identification and dating of sequences, sequence boundaries, and systems tracts,
which we can then evaluate in terms of lithology and reservoir potential. The multi-step
interpretation procedure outlined in Table 1 (below) is modified from that of Vail and Wornardt
(1991), and can be applied at both the regional or prospect level.

In outlining this methodology, our goal is to show how it may be applied, and what to look for in
the interpretation process. While the methodology outlined in Table 1 is linear; it requires
constant feedback and reworking of previous steps to obtain satisfactory results. Seismic
sequence stratigraphic interpretations require an artistic, "right-brain" flair combined with a firm
understanding of sequence stratigraphic concepts. It should be emphasized that only one correct
interpretation for a given geologic setting exists (Vail, pers. comm.), and that it may require
numerous guesses and re-evaluations to obtain this absolute truth when studying the distant
geologic past. While this methodology presents useful guidelines, it should serve primarily as a
basis for the development of our own procedures, based upon experience, type location, and
available tools and data.

Table 1. An integrated sequence stratigraphic interpretation procedure. (Modified from


Vail and Wornardt, 1991)

STEP PROCEDURE TOOLS


1 Structural interpretation of seismic data Seismic profiles
2 Interpret lithologies from well log character Well logs
3 Interpret depositional environments and paleobathymetry Well logs & paleo data
4 Identify major 2nd order sequences Seismic profiles & well logs
Faunal abundancies and
5 Interpret condensed sections
diversities
6 Establish condensed section ages Biostratigraphic chart
7 Locate discontinuities Dipmeter log
8 Interpret SBs & systems tract boundaries Well logs
Seismic profiles & tie to well-
9 Interpret SBs, mfs and systems tracts
logs
10 Tie interpreted features to sea level cycle chart Global or local cycle chart
11 Identify & coordinate parasequences and marker beds Well logs
12 Construct seismic sequence stratigraphic cross sections Well logs and seismic
Well logs biostratigraphy&
13 Produce chronostratigraphic display summary
seismic

Seismic Profiles
The first step in deciphering sequence stratigraphy from seismic reflection profiles is to interpret
the structure, major faults, and seismic reflector terminations and truncations of the area. We do
this using the standard loop-tying procedure across the seismic grid. We usually start in the least
complex areas and work toward more complex areas.

As we learned earlier; the geometries of seismic reflectors are the key to deciphering sequence
boundaries and their intervening sequences. To extract sequence bounding surfaces, we mark
reflector terminations representing the downlap, onlap and offlap stratigraphic surfaces. We do
this using small arrows at critical points, as illustrated on the seismic profile of two original seismic
stratigraphic models: Figure 1 (a generalized stratigraphic section of a sequence, with the
vertical axis measured in depth) and Figure 2 (a generalized chronostratigraphic section of a
sequence, with the vertical axis measured in time ).

The overall morphologies revealed by the sequence boundaries on dip sections provide a
general indication of the former margins of a basin, as well as its paleobathymetric environments
and sequence stratigraphy. Tracing the offlap break inflection points, which separate relatively flat
topsets from dipping foresets, may help identify the location and evolution of the shelf break of a
basin margin through time.
At this point, insight from an appropriate model, such as the one illustrated in Figure 3 (Log
responses of different systems tract on the shelf and slope), should help us identify stratigraphic
sequences, while internal morphologies and paleo water depths should give us clues as to the
types of systems tracts present, or at least expected. The procedure of "pushing"
appropriately coded colors along sequence boundaries, and within identified systems tracts,
expedites the transfer of interpretations to base maps, the contouring of time values, and the
production of cross sections. Converting time values to depth values is helpful in detecting
velocity-induced false structures.

The regional seismic stratigraphic interpretation ( Figure 4 ) is an excellent example of what an


interpretation should look like. At this point, the input of well-log and high-resolution
biostratigraphic data is crucial towards developing the investigation beyond the regional level.
Well Logs and Cores
The integration of well log data, including SP, gamma ray and resistivity curves, sonic log
velocities, and core samples analyses, enables us to make a reasonably accurate prediction of
lithologies and fluids. We can use well data to identify sequences and systems tracts, and to date
condensed sections and sequence boundaries.

The detailed interpretation of well log electric curves consists of first verifying and picking the
sequence boundaries, which we correlate with the seismic reflection profiles at the well site using
synthetic seismograms. A synthetic seismogram is a continuous, vertical record that looks like a
seismic trace but is derived from sonic and density logs (Risch et al., 1994a). Synthetically
derived seismic traces are the best tool for correlating between time-based seismic reflection
profiles and depth-based well data, and are essential for obtaining an optimal interpretation. Once
we have bridged the gap between seismic and well log data, we can then proceed with the
detailed analysis and interpretation of sequence stratigraphy from well log data.

Figure 1 (Characteristic well-log responses and associated lithologies of the basin floor fan),
Figure 2 (Characteristic well-log responses and associated lithologies of the slope fan
complex)

Figure 3 (Characteristic well-log responses and associated lithologies of the prograding complex)
show the idealized electric well log traces. In particular, we may note the following:

• Basin floor fan (Figure 1 ): This unit has a characteristic blocky, "boxcar" character. The
interval itself typically consists of a massive sandbody, although thim pelagic shales can
occur interbedded with the sands. At the outer edges If the sandbody, interbedded shales
may become common and form local barriers to vertical flow. Individual sand layers
typically coalesce in the central sandbody, resulting in a single hydrocarbon-water
contact.

At the upper boundary, pelagic shales may result in an excellent seal; if channel
overbank sands of slope fan rest directly on a basin floor fan, however, they form a poor
seal. The top of the basin fan typically has a thin transition zone, but an abrupt top may
indicate postdepositional submarine current erosion.

At the sequence boundary, the typically massive sand rests directly on pelagic shales or
marls. Biodirectional downlap may be recognized from log correlation. Erosion on the
boundary is uncommon; when it does occur, it can indicate proximity to the sluiceway,
down which sands are transported to the central basin.

• Slope fan complex (Figure 2 ): The "nervous" log character signifies the beginning of the
upper boundary, and indicates a shift from overlying pelagic shale to laminated sand-
shale of the slope fan. The overlying shale typically has a faunal abundance peak.

Within the interval, the channel-overbank units show a crescent shaped log character.
The upper part of the unit consists of overbank sands that thin and become sparse
upward, while the lower part consists of attached-lobe sands that thicken and become
more abundant upward. Wells in the central lobe may show coarsening-upward channel
sands of "multi-story" highly sandy intervals. Sub-parallel facies are poorly understood,
but may consist of sheet-like attached lobe facies in areas of abundant sand supply. Six
to ten channel overbank units may stack in one lowlands systems tract.

The lower boundary lies on the sequence boundary or on a basin floor fan. Faunal
abundance peaks may separate units and may seal the basin floor fan.
• Prograding complex (Figure 3 ): The top of the prograding complex marks the transition
from upward-shallowing to upward-deepening sediments. Toplap is common below the
boundary, and there may be a transgressive surface of erosion at the boundary.

Thick intervals of coarsening-upward sands may occur near the top of the sand. Thinner,
blocky sands may underlie the boundary. Near the top of the interval, shoreline and
deltaic sands occur, which grade downward into prodelta and pelagic shales. The
prograding complex pinches out against the offlap break (shelf edge) of the previous
highstand systems tract, while shingled turbidite sands may occur at the base.

At the lower boundary, clinoform toe pelagic shales rest on "nervous" laminated silts,
sands and shales of the slope fan complex. A significant faunal abundance peak
commonly in the basil pelagic shales. Note that the displayed electric log curves are
idealized and that actual well log data will consist of an infinite number of variations on
these themes. Variations are frequently indicative of the changing nature of depositional
environments through time, which is precisely the kind of information we should be
looking for in our search for hydrocarbon traps.

Thus, the slope fan complex illustrated in Figure 2 may consist of repetitions of the same
pattern and indicate the presence of several repeated or "stacked" slope fan complexes
— an excellent potential hydrocarbon prospect. Or; the well-log data might show a lack of
high-amplitude "kicks" on the electric or gamma ray logs, indicating low sand contents
within an sfc unit a poor exploration prospect.

The power of the sequence stratigraphic method, however, allows us to understand the
basic sedimentary depositional processes and environments at a given subsurface point
and to extrapolate these laterally in the search for optimal hydrocarbon prospects. The
characteristic well log patterns provide us with a rough guide for interpreting sequence
stratigraphic units from either SP or gamma ray well logs.

We should reinforce and confirm our interpretations using synthetic seismograms. Figure
4 shows a correlation between a well log curve, synthetic seismic profile, and a seismic
reflection record. Note that the area below the label "LOCAL TOP" displays the
characteristic well log pattern indicative of a basin floor complex and overlying slope fan
complex. Synthetic seismic traces and electric logs, generated at the same scale, can be
placed side-by-side for better correlation and for comparisons to additional data, such as
high-resolution biostratigraphy and paleo water depths. We must continuously cross-
correlate between the various tools, refining our interpretation in an iterative process. It is
convenient to have all data on the same scale; for example, 1 inch = 1,000 feet.

The interpretation of lithologies and associated systems tracts from well logs involves
matching characteristic signatures . In Figure 5 (Characteristic well-log responses and
associated lithologies of the transgressive systems tract ) and Figure 6 (Characteristic
well-log responses and associated lithologies of the highstand systems tract ), we see
idealized gamma ray log curves through a transgressive and a highstand systems tract,
respectively.

• Transgressive systems tract ( Figure 5 ): The lowest resistivity and highest gamma ray
values indicate the most clay-rich shale at the maximum flooding surface. A faunal
abundance peak is common at this boundary. Also, there is a discontinuity surface,
indicating downlap above and apparent truncation below.

The interval itself becomes finer-grained and thinner-bedded upward. Neritic shales form
the best seal near the top, while estuarine, beach and shoreface sands occur near the
base. The basinal equivalent is a pelagic shale. Ravinement surfaces backstep on
successive parasequence boundaries and may underlie beach-shoreface sand bodies,
while barrier islands and shoreface sands can be well-preserved and more permeable
than highstand systems tract sands. Lowstand incised-valley erosion is common at the
sequence boundary; below the boundary, erosional truncation is common

• Highstand systems tract ( Figure 6 ): Lowstand erosion and incised valleys are common
at the sequence boundary, while submarine canyons cut into the shelf of the preceding
highstand systems tract. Truncation and toplap are common below the boundary.
Shoreline and deltaic sands predominate near the top of the interval; sands become
coarser-grained and thicker-bedded upward. The interval progrades into neritic shales
toward the basin, and the basinal equivalent is pelagic shale. Correlation of sands is
difficult due to discontinuous coastal plain and alluvial facies; consequently, reservoir
continuity is only fair to poor.

Downlap of shale-rich clinoform toes onto the boundary of the maximum flooding surface.
There may also be a planktonic abundance peak characteristic of a thin limestone.

We continue our analysis, using an iterative process between seismic profiles, synthetic
seismic traces and well log data. We perform this back-and-forth procedure until we
obtain a satisfactory interpretation, have identified all sequences and systems tracts in a
given well, and have tied the interpretation to the seismic data (Figure 7, Integrated
sequence stratigraphy of seismic profile with well log profile and correlation, Gulf of
Mexico).

The interpretation shown in Figure 7[atxht]0 results from the integration of biostratigraphic data, from
which condensed section ages and paleo water depths have been extracted, and correlation with the gamma
ray log to the right. The integrated sequence stratigraphic methodology provides a continuous history of
sedimentary deposition along the seismic profile, enabling us to extract systems tracts and to evaluate their
hydrocarbon potential.
Minor, repetitive incursions of well log curves within individual systems tracts are indicative of
parasequences, which we can best identify from continuous core samples. While using cores and
cuttings to confirm our lithologic interpretations of well data is highly recommended, the use of
high-resolution biostratigraphy is a must.

High-resolution Biostratigraphy
Biostratigraphy is the study of the paleontological aspects of rocks, and the differentiation of rock
units through the study of the fossils they contain (American Geological Institute, 1984).
Biostratigraphy involves extracting, cataloguing and determining the species of microscopic
fossils in the sedimentary record. Different species, such as benthic bottom dwellers or pelagic
surface dwellers, may indicate particular environments and ecological habitats.

The input of biostratigraphic information is essential to a successful seismic sequence


stratigraphic interpretation and reduces the risks in exploring for hydrocarbons. Biostratigraphic
data is useful not only for dating purposes, but also for determining past depositional
environments and associated paleobathymetry, since depositional environments enable us to
understand the sedimentary record in terms of sequence stratigraphy.

High-resolution biostratigraphy consists of detailed paleontologic sampling at regular downhole


intervals of 30 to 60 feet (Vail and Wornardt, 1991). We can then tabulate a paleontological
checklist of fossil species content and abundances, showing quantitative variations of fossil
species with depth, which we can then correlate to well logs and seismic sections.

Such a checklist provides a simple visual display of species content and abundances, relative to
well depth. Biostratigraphic results displayed in this manner represent a statistical approach,
which provides a practical method for extracting biostratigraphic and chronostratigraphic data. In
addition, this statistical approach represents a reliable and repeatable method for identifying
paleoenvironments and for selecting condensed sections. Condensed sections are characterized
by an abundance of fossils associated with periods of low terrigenous sediment input, when
sedimentation is restricted to pelagic and hemipelagic deposition.
This is clearly shown by the microfossil histograms, and interpreted systems tracts of Figure 1
(Microfossil content versus well depth). This slow sedimentation results from the removal of
sedimentary sources and deepening marine conditions. Since condensed sections are associated
with the maximum flooding stages and coastal onlap, their detection is essential in the
interpretation of stratigraphic sequences.

These histograms enable us to recognize important biostratigraphic references from the patterns
produced by fossil abundances and diversities. We can readily see the presence of a maximum
flooding surface and a condensed section between well depths 2640 and 2790 feet on Figure 1 .
Note the annotated abundance and diversity values and the selection of systems tracts
boundaries.

When integrating biostratigraphic data with well logs and seismic reflections, we can generally
locate sequence boundaries directly above fossil abundance peaks for deepwater environments.
In addition, the telltale electric log readings and the nature of seismic reflectors combine to
provide us with a high level of confidence as we choose a particular sequence boundary or
subunit.

Biostratigraphic data also reveals the paleobathymetry of stratigraphic formations, as certain


fossils, particularly benthic bottom-dwellers, are associated with well-defined marine
environments Associated with water depth. Hence a plot of paleo water depths, derived from
biostratigraphic data, overlain on a graph of well log and paleo data, enables us to confirm our
interpreted depositional environments. This process reduces such errors as interpreting a blocky
sand to represent a basin floor fan when paleo water depths indicate the presence of an inner
neritic environment.

The assembly of an integrated graphical data display, including all of the above tools,
tremendously enhances our confidence in deciphering the sequence stratigraphic record. Once
we assemble an integrated graphical display, we should then correlate it to an appropriate sea
level cycle chart to further reinforce our interpreted depositional environments, ages, and
associated lithologies.

At this point in our integrated sequence stratigraphic interpretation, the emerging picture of the
subsurface should enable us to predict the lateral and vertical distribution of source rocks,
reservoir sands, and seals. The precision of our predictions is a function of the well data and
seismic grid spacing more than anything else. For exploration at the prospect level, 3-D seismic
coverage can provide the requisite resolution for detecting small petroleum traps, while a more
widely spaced seismic grid will provide only a general idea of potential targets.

Sea Level Cycle Charts


We can greatly enhance our seismic sequence stratigraphic interpretations by tying them to a sea
level chart, both to verify the interpretation and to provide a chronostratigraphic reference. We do
this by comparing our interpretations to either the global cycle chart, from Haq et al. (1987), or an
appropriate locally derived chart

Because the global cycle chart of Haq et al. covers a long period of geologic history, it is
consequently limited to the extraction and correlation of first- through third-order stratigraphic
sequences. We may need a more detailed chart when seeking finer details at the prospect and
production level, when stratigraphic sequences of the third-order and smaller are involved. With
time and experience, individual exploration teams develop their own charts, based upon their
preference of chronostratigraphic events, and how these fit into a particular geologic setting. The
global cycle chart serves more frequently as a general guideline until further studies refine it into
smaller time periods for specific sedimentary basins.
With the accumulation of experience from seismic sequence stratigraphy in exploration efforts
around the world, minor modifications to current sea level cycle charts are inevitable, as the
science is still in its infancy. As work progresses, sequence stratigraphic models for various
settings will be refined.

Prediction of Lithofacies and Depositional Environments


A completed sequence stratigraphic interpretation is a powerful tool for predicting subsurface rock
properties. A review of the integrated data, assembled from wells and seismic reflection profiles
and displayed as stratigraphic charts, can reveal depositional environments and associated
lithologies.

Using displays of this type, we can confidently assess potential producing zones. We can target
particular zones for exploration, and analyze these further in terms of source rocks, seals and
potential reservoir traps. Understanding the sequence stratigraphy of an area not only enhances
our confidence in an interpretation and lowers exploration risks, but most importantly provides an
accurate framework for laterally extrapolating depositional environments and lithologies away
from a well-site. The modeling of seismic reflection character for potential stratigraphic traps,
based upon detected lithologies, provides additional confirmation and verification of targeted
hydrocarbon traps. As a general rule of thumb, we assume that long-term rises in relative sea
level, such as those associated with second-order tectonic cycles, tend to produce source rocks,
while the falling phase of these long-term cycles produce the reservoir rocks (Vail, pers. comm.)

Reservoir Exploitation and Production


Applying sequence stratigraphy, using the procedures outlined above, enables the seismic
sequence stratigrapher to make valid, and potentially economic, predictions as to the composition
of sedimentary rocks in the geologic record.

To obtain meaningful results, we must do more than follow a simple procedure; we must develop
a thorough understanding of the sequence stratigraphic concepts and hone our interpretation
skills through practical experience. By using all available data and our best abilities in our quest
for hydrocarbon resources, we can make more than an educated guess about the nature of the
subsurface geology at both the regional and prospect levels. Sequence stratigraphy provides a
comprehensive framework not only for understanding cyclic depositional processes but more
importantly for detecting the resulting sedimentary deposits in the subsurface.
The mapping of reservoir-quality sands and seals, together with an understanding of the source
rocks and migration paths, enables us to select high-quality prospects, such as those associated
with the lowstand, transgressive and high-stand systems tracts ( Figure 1 , Siliciclastic sequence
stratigraphic model showing systems tracts and the distribution of sands). Reservoir-quality sands
are also distributed throughout most of the other systems tracts.
6. CONCLSUINS AND CURRENT CONTROVERSIES

The Model
The idea of the sequence as a basic depositional unit produced by an oscillation in sea level has
rarely been in contention. But, there has been some controversy as to the definition of sequence
boundaries. For example, the initial genetic model of Galloway (1989a) generated a great deal of
confusion until it was harmonized with the Vailian model (Xue and Galloway, 1993). The choice of
sequence boundaries is essentially a personal preference, since it makes no difference where we
start and end a cyclic process.

The sequence stratigraphic models are just that, models. We should consider them examples
culled from extensive observations. These models provide guidelines, and should not be taken as
firm templates for every conceivable depositional situation, as every stratigraphic sequence will
be unique. In addition, sequence boundaries and systems tracts all have diagnostic, but not
unique, seismic and well log signatures. Blind adherence to a model will necessarily reveal
certain "flaws" when applied to a real-world depositional system. We should scrutinize the
message of these "flaws" and use them to improve our application of the method, rather than to
dismiss the method itself.

Nomenclature
Nomenclature, while sometimes confusing, nevertheless serves an important purpose, since it is
essential that investigative results be clearly communicated through the use of standard
terminology. These terms may not be precise or semantically correct, but they must serve the
purpose of effectively conveying a point or description. The important issue is that as long as the
terms are defined and understood, they will help advance the discipline of modern stratigraphy
with a minimum amount of confusion.

Many arguments against sequence stratigraphy arise from the lack to date of formal acceptance
of sequence stratigraphic terminology into the code of stratigraphic nomenclature;... we
nonetheless strongly believe that the geologic community, rather than a stratigraphic commission,
will determine which terminology ultimately becomes adopted; and regardless of the
"correctness" of sequence terminology, the geologic community at large will decide on the basis
of which terms they routinely use, and which terms they do not, those terms that will stand the
test of time. (Weimer and Posamentier, 1993)

Figure 1 lists the terminology, color codes, and systems tract graphic demarcation lines currently
used by Peter Vail and colleagues at Rice University, and to a certain extent within the petroleum
industry today. Use of these standards will ease the interpretation process, particularly when
dealing with large volumes of seismic and well data, as well as the presentation and
communication of results.
Sea Level Curves and Global Eustasy
Upon publication of AAPG Memoir 26, the global nature of the sea level cycle chart was
immediately challenged, since Exxon's research team could not reveal the proprietary data used
to develop the chart. This issue was partly put to rest when Haq et al. (1987) published the
location of outcrops used in its compilation. Nevertheless the global sea level cycle chart
continues to foment criticism (Miall, 1986, 1991, 1992), in spite of more than twenty successful
years of application worldwide. While the precise causes and timing of sea level cyclicity are still
poorly understood, this should by no means prevent its application, as advocated by Miall (1986),
and Hubbard (1988) since the Exxon sea level curves were based on numerous well-founded
observations. Miall's (1992) statement that "the existing Exxon cycle chart should be abandoned
— it is too flawed to be fixed.. .we should start again, by building a framework of independent
sequence stratotypes. ..without preconceptions as to the results" would require reinventing the
wheel, considering the enormous amount of research that went into developing this chart. There
was a very strong basis for the initial global sea level curve, which is constantly being revised as
sequence stratigraphers around the world apply it to their regional basins (Wornardt and Vail,
1991). Some workers who initially expressed skepticism regarding its validity have changed their
views through application of the concept. Frank Brown (1993), a preeminent explorationist and
teacher of seismic and sequence stratigraphy, describes his initial experiences with the subject:

Much of my earlier concern about and disagreement with Exxon's global cyclic sea-level concepts
probably sprang from limited data quality and my own early lack of an intensive experience
focused on subtle reflection anomalies. ... At this point in time (1993)... despite the many short
courses and debates over the matter, most geoscientists are still just talking, typically defending
their previous ideas and rarely sitting down and undertaking a meaningful, data-intensive effort to
evaluate objectively these new ideas.

The published global sea level cyclicity curve was, however based primarily upon exploration
seismic data, which limited resolution to first- through third-order sequences, with anything
smaller falling into the realm of the parasequence at that time. With further studies and
observations, however the sea level curve is being refined into more numerous and smaller
sequence orders, at least within well-studied basins (e.g., Brink et al., 1993). In addition, we
should note that the precision with which chronologic measurements of rocks are determined still
leaves much to be desired. With improved dating techniques, the accuracy of correlating
stratigraphic sequences on the global cycle chart will be enhanced (refer to Berggren et al.,
1996).

There are a few resolvable problems within the greater issue of an accurate chronostratigraphy:
first, an appropriate nomenclature for individual sequences, and second, how these sequences
should be defined from the biostratigraphic record. Exxon Production Research originally
assigned numbers defining each sequence on the basis of its age, a straightforward
methodology. This method allows additional sequences to be inserted as they are discovered.
The biostratigraphic designation of sequences poses a minor quandary to the proper naming of
sequences with respect to biomarkers. If a sequence boundary is located just above a species
extinction, can it rightly be named after that particular species, or should the extinction designate
the underlying (earlier) sequence boundary? These questions should eventually be resolved per
the comments of Weimer and Posamentier quoted above.

A subject of much debate, however; has been the causes of eustasy. The orbital-climatic
Milankovitch effects were not assessed in detail when the concept was first made public, and are
still poorly appreciated by many. However; the field of cyclostratigraphy has emerged from the
study of sequence stratigraphy, and will more than likely provide new insights into the issue as
time passes. As to whether the causes of relative sea level changes are tectonic or eustatic, it is
clear that relative sea level changes represent the summation of a number of natural phenomena.
Confusion over the meaning of the terms eustasy and relative sea level have led certain
geoscientists astray (Weimer and Posamentier; 1993). To quote further from Weimer and
Posamentier (1993), "... regardless of whether one accepts or rejects the published global sea
level curves, it is important to realize that this debate does not affect the other major; and much
more important, aspect of sequence stratigraphy: i.e., lithology prediction."

In summary, we should first apply the method, and then come to our own conclusions concerning
the validity of this aspect of the sequence stratigraphic concept.

The Question of Data Bases


Many subjects depend upon our perspective, and seismic sequence stratigraphy is no exception.
The sequence stratigrapher who is limited to just a single tool, such as 2-D seismic data, will have
great difficulty in applying the concept compared to someone equipped with all the tools
necessary for a thorough, integrated approach. We must look at the geological column using a
number of tools to develop the most complete depositional history. To do this, we must assemble
data bases from several disciplines, which may require input from explorationists trained in many
disciplines, or the development of a multidisciplinary team, as is currently practiced within the
petroleum industry.

We must consider the "big picture" before pursuing exploration efforts at the prospect scale, or
we will not develop a comprehensive picture. Being specialized in one discipline, such as well
logging or petrophysics, will not enable us to effectively extract a complex set of past geologic
processes from the sedimentary record. Only through a methodical approach, starting with
seismic stratigraphy and then gradually including well data, biostratigraphy, and synthetic
seismograms, can we obtain a complete picture. Hence, some of the criticism levied against
sequence stratigraphy has been based on the use, or attempted use, of limited data bases. This
is a narrow point of view for a relatively complex concept.
GLOSSARY
accommodation space: the space made available for potential sediment accumulation ... which
... is a function of both sea level fluctuation and subsidence (Jervey, 1988). Accommodation
space can be positive, implying deposition, or negative, implying erosion and the formation of a
sequence boundary. Variations in shelfal accommodation space control the sequence
development geometry, i.e., whether prograding, aggrading, or retrograding.

aggradation: accumulation of a facies or group of facies within a depositional system, with no


significant lateral migration of the facies or facies belts towards either the land or basin. Occurs
when the increase in accommodation space is equal to the sediment supply rate. Causes
parasequences to stack vertically. Requires a perfect balance between sediment supply rate and
rise in relative sea level (Van Wagoner et al, 1990).

apparent truncation: describes the termination of relatively flat-lying seismic reflectors beneath a
dipping seismic surface, where that surface represents a marine condensed section. May look
like erosional truncation, but represents a facies change from appreciable sedimentation to
condensed section. Frequently characterizes the top surface of the transgressive systems tract.

base level: a) the theoretical limit or lowest level toward which erosion of Earth's surface
constantly progresses but seldom, if ever, reaches; esp. the level below which a stream cannot
erode its bed. The general or ultimate base level for the land surface is sea level; b) a curved or
planar surface extending inland from sea level, inclined gently upwards from the sea and
representing the theoretical limit of stream erosion; c) the surface toward which external forces
strive and at which neither erosion nor deposition takes place (Barrel, 1917); a surface of
equilibrium (American Geological Institute, 1984). In sequence stratigraphy, a change in base
level associated with falling sea level leads to erosion and progradation, while rising sea level
leads to accumulation and retrograde deposition. Sediments are eroded above a base level and
aggrade below it.

biostratigraphy: the study of the paleontological aspects of rocks, and the differentiation of rock
units through the study of the fossils they contain (American Geological Institute, 1984). High-
resolution biostratigraphy is performed on wells at small sampling intervals (e.g., 30 to 60 feet).

bright spot: seismic reflection amplitude anomaly caused by the high reflection coefficients
(sometimes) associated with the presence of gas.

clinoform: a sloping depositional surface, commonly associated with strata pro-grading into deep
water (Mitchum, 1977). Categorized according to shape: sigmoidal, oblique, tangential, parallel,
complex and shingled, and containing bottomsets, foresets and topsets. Clinoforms record the
migration of a sloping sediment surface. In a basin margin setting, clinoforms generally record the
progradation of a shelf-slope system into deeper water.

coastal onlap: onlap is a base-discordant relation in which horizontally inclined strata terminate
progressively updip against a surface of greater initial dip. Coastal onlap is the progressive
landward onlap of the coastal (littoral or coastal non-marine) deposits in a given stratigraphic unit.
The vertical and horizontal components of coastal onlap are coastal aggradation and coastal
encroachment. Coastal onlap is generally inferred from seismic data as the landward onlap of
topset reflectors. Does not necessarily occur at the coastline (recent usage by Exxon extends this
concept landward), nor need it involve coastline facies, i.e., coastal onlap could represent
lagoonal facies.

composite sequence: a succession of genetically related sequences in which the individual


sequences stack into lowstand, transgressive and highstand sequence sets. (Van Wagoner et al.,
1990). Involves the stacking of higher order sequences (i.e., third order) within lower order
sequences (i.e., second order).

condensed section: a thin, marine stratigraphic unit characterized by very low sedimentation
rates and consisting of pelagic to hemipelagic sediments. Condensed sections reach their
maximum aerial extension at the time of maximum regional transgression of the shoreline, and
are diachronous across a basin. (Loutit et al, 1990). As the product of very low rates of
sedimentation, a condensed section may be associated with hardgrounds, glauconitic bands, bio-
erosion, submarine winnowing, and faunal abundance peaks. Have a characteristic signature on
well logs- a thin unit of anomalously high gamma, high or low sonic velocity, and/or high density.
Biostratigraphic analysis is a powerful tool for identifying condensed section intervals.

continental shelf: the part of the continental margin between the shoreline and the continental
slope; or, when there is no noticeable slope, a depth of 200 m. Characterized by a very gentle
slope of 0.1 degrees. (American Geological Institute, 1984).

continental slope: the part of the continental margin between the continental shelf and the
continental rise or oceanic trench. Characterized by a relatively steep slope of 3 to 6 degrees.
(American Geological Institute, 1984).

correlative conformity: the correlative conformity to a sequence boundary is the depositional


surface at the base of the lowstand systems tract (or shelf margin systems tract).

cyclothem: repetitive strata resulting from repeated marine transgressions and regressions
(Wanless and Weller, 1932).

depositional sequence analysis: an integrated method for conducting a seismic sequence


stratigraphic interpretation. Involves a number of tools, including seismic reflection profiles,
synthetic seismograms, vertical seismic profiles, well logs and biostratigraphy.

depositional system: a three-dimensional assemblage of lithofacies, genetically linked by active


(modern) or inferred (ancient) processes and environments (Fisher & McGowan, 1967; Brown &
Fisher, 1977). Grouping of lithofacies into common workable groups, such as submarine fan
systems, barrier island systems, and alluvial systems. Recognized on logs, cores, and seismic
profiles as an interval of related facies assemblages. For example, depositional systems
associated with the highstand systems tract would consist of alluvial, coastal plain, shallow
marine and slope to deepwater environments.

diachronous: literally, "across time." Applied to lithological units (e.g., a sandstone bed, a
limestone reef) which appear to be continuous beds but which in fact represent the development
of the same facies at different places at different times. Said of a rock unit that is of varying age in
different areas or that cuts across time planes or biozones; e. g. , a marine sand that was formed
during an advance of a shore line (transgression) and becomes younger in the direction in which
the sea was moving. (AGI Dictionary of Geological Terms, 1984).

downlap: a base-discordant relation in which initially inclined strata terminate downdip against an
initially horizontal or inclined surface (Mitchum, 1977).

equilibrium point: the point along a continental shelf profile where the rate of eustatic change
equals the rate of subsidence. Seaward of this point, the rate of subsidence is greater than the
rate of eustatic fall, resulting in the addition of new space; landward of this point, the opposite
occurs. Alternatively, the equilibrium point defines two zones: 1) a zone of relative sea-level rise
seaward of the equilibrium point and, 2) a zone of relative sea-level fall landward of the
equilibrium point (Posamentier and Vail, 1988).
eustasy: a change in elevation in sea level on a worldwide basis relative to a stationary datum
such as the center of Earth (Christopher et al., 1988). Eustatic sea level is a component of
relative sea level and is independent of uplift or subsidence of the land surface. Eustasy is
recognized from global correlations of relative sea level change. Eustasy occurs on a wide variety
of scales, termed "orders" in sequence stratigraphy.

genetic sequence or unit: a package of sediments recording a significant episode of basin


margin outbuilding and basin filling, bounded by periods of widespread basin-margin flooding
(Galloway, 1989a). Now accommodated with the clastic sequence model of Vail (Xue and
Galloway, 1993).

hiatus: a) a break or interruption in the continuity of the geological record, such as the absence in
a stratigraphic section of rocks that would normally be present but are missing either because
they were never deposited or because they were eroded before the deposition of the beds
immediately overlying the break; b) a lapse in time, such as the time interval not recorded by
rocks at an unconformity. The term has been used for the time value of an episode of
nondeposition and for the time value of nondeposition and erosion together (American Geological
Institute, 1984).

highstand: interval of time during a cycle or cycles of relative change of sea level when sea level
is above the shelf edge in a given local area (Mitchum, 1977).

highstand systems tract: the upper systems tract in either a Type 1 or a Type 2 sequence (Van
Wagoner et al., 1988). A unit of rock bounded below by a maximum flooding surface and above
by a Type 1 sequence boundary. The highstand systems tract is commonly widespread on the
shelf and may be characterized by one or more aggradational parasequence sets succeeded by
one or more progradational parasequence sets with prograding clinoform geometries.
Parasequences within the highstand systems tract onlap onto the top of the transgressive or
lowstand systems tracts in a landward direction and downlap onto the top of the transgressive or
lowstand systems tracts in a basinward direction.

HCI: hydrocarbon indicator, usually by means of remote sensing, such as geophysical reflection
profiles.

incised valley: entrenched fluvial systems extending their channels basinward and eroding into
the underlying strata in response to a relative fall in sea level. On the shelf, the incised valleys are
bounded below by the sequence boundary and above by the first major marine flooding surface,
called the transgressive surface. Can be part of either transgressive or regressive relative sea
level phases.

inflection point: the points on a relative sea level curve of greatest rate of rise and fall of relative
sea level. The sea level curve is usually depicted as a sine wave, where inflection points are
apparent. The curve may, however, be depicted as a saw tooth, with inflection points taken when
and where sea level changes occur, so the change from falling to rising phases may be sudden.

insolation curve: curve depicting the variation of incoming solar radiation to Earth, usually for a
fixed latitude. Originally derived by Milankovitch (1920) based upon variations in Earth's orbit and
inclination relative to the sun.

lowstand: interval of time during a cycle or cycles of relative change of sea level when sea level
is below the shelf edge. A comparative lowstand occurs where sea level is at its lowest point on
the shelf during the deposition of a series of sequences (Mitchum, 1977). Eustatic lowstands are
times of generally low relative sea level. The EPR model predicts that a distinctively evolving set
of facies, the lowstand systems tract, will be associated with sea level lowstands.
lowstand systems tract: the lowermost systems tract in a depositional sequence, if it lies
directly on a Type 1 sequence boundary (Van Wagoner et al., 1990). The lowstand systems tract,
if deposited in a basin with a shelf break, can generally be subdivided into three separate units —
a basin floor fan, a slope fan, and a lowstand wedge (prograding complex).

marine flooding surface: surface separating younger from older strata across which there is
evidence of an abrupt increase in water depth. This deepening is commonly accompanied by
minor submarine erosion or non-deposition (but not by subaerial erosion due to stream
rejuvenation or a basinward shift in facies), with a minor hiatus indicated. The marine flooding
surface has a correlative surface in the coastal plain and on the shelf (Van Wagoner et al., 1990).

marine onlap: the onlap of marine strata (Mitchum, 1977). Recognized on seismic data by the
termination of low-angle reflectors. Marine onlap should not be confused with coastal onlap,
which occurs beyond and below the offlap break (shelf edge), is unaffected by, and cannot be
used to determine, relative sea level.

maximum flooding surface: the surface corresponding to the time of maximum flooding; also
called the downlap surface (Posamentier et al, 1988). The most landward flooding surface within
a transgressive/regressive cycle. Commonly associated with a condensed section seaward, and
forms the boundary between the transgressive and highstand systems tracts. Represents the
downlap surface of the highstand systems tract on seismic profiles. Also represents the bounding
hiatal surfaces that defined the genetic stratigraphic units of Galloway (Xue and Galloway, 1993).

maximum progradation surface: the conceptual surface forming the boundary between a
progradational unit and the overlying retrogradational unit (Milton, 1990). Defines the sequence
profile at the time that the basin margin facies belts were at their most distal position into the
basin — the point at which seismic reflectors begin to retrograde.

megasequence: sequence sets where prominent regional unconformities separate the total
sedimentary section into major, discrete phases of basin evolution. The boundaries between
these major packages record the principal periods of change in basin drive mechanism,
geometry, and polarity (Hubbard, 1988). Megasequences are controlled by tectonic processes
and are not synonymous with second-order sequences. Instead, they represent distinct tectonic
phases such as pre-rift, syn-rift and post-rift periods.

offlap break: in a basin-margin setting, the inflection point in seismic reflectors marking the
boundary between relatively flat, planar topsets and clinoforms. Believed to form in the vicinity of
the fair-weather wavebase (Vail et al, 1990). Recognized on seismic data as the point of change
in slope between topset and foreset clinoform reflectors, and may coincide with the shelf break.
Marks the boundary between shelf and slope processes.

order (of cyclicity): three major orders of cycles are superimposed on the sea-level curve.
Cycles of first, second and third orders have durations of 200-300 million, 10 to 80 million and 1
to 10 million years respectively (Vail et al, 1977). Development of a global sea level chart based
upon sequence analysis reveals a hierarchy of relative changes in sea level, with shorter duration
cycles nestled into longer duration cycles. Orders have been assigned to cycles of set time
lengths and ranges as per chart below (Vail et al, 1990).
Order Duration (million years)

1 >50

2 5 to 50

3 .5 to 5

4 0.1 to 0.5

5 0.01 to 0.1

6 <0.01

parasequence: a relatively conformable succession of genetically related beds or bed sets


bounded by marine flooding surfaces and their correlative surfaces. In special positions within the
sequence, parasequences may be bounded either above or below by sequence boundaries (Van
Wagoner et al., 1990). Parasequence boundaries form the lowest-order practical
chronostratigraphic unit for correlation and mapping. Parasequences are the building blocks of
sequences, with characteristic geometries indicating progradation, aggradation and
retrogradation, and are believed to be controlled mainly by sediment supply rate.

parasequence set: a succession of genetically related parasequences forming a distinctive


stacking pattern bounded by major marine-flooding surfaces and their correlative conformities
(Van Wagoner et al, 1990).

pro gradation: the basinward migration of a facies belt within a depositional system. Refers to
the advance of facies or depositional systems into a basin, and is indicated by clinoforms on
seismic data.

ravinement surface: the disconformity caused by passage of the surf zone of a transgressive
sea, which may bevel the marginal deposits of the coast being transgressed (Stamp, 1922). The
term is sometimes used to include the entire transgressive erosional surface below the
transgressive systems tract.

sequence: a relatively conformable success ion of genetically related strata bounded above and
below by unconformities or their correlative conformities (Mitchum et al., 1977).

sequence stratigraphy: the study of rock relationships within a chronostratigraphic framework of


repetitive, genetically related strata bounded by surfaces of erosion or nondeposition, or their
correlative conformities. (Van Wagoner et al., 1988).

shelf break: the edge of a continental shelf where the slope increases appreciably.

synchronous: existing or occurring at the same time; contemporary or simultaneous. The term is
applied to rock surfaces on which every point has the same geologic age (AGI Dictionary of
Geological Terms, 1984).
RECOMMENDED READING LIST
*Berg, O.R. and D.G Wolverton, eds., 1985. Seismic stratigraphy II-An integrated approach to
hydrocarbon exploration. AAPG Memoir 39. Tulsa, OK: AAPG.

Cross, T. A., ed., 1990. Quantitative dynamic stratigraphy. Englewood Cliffs, N.J.: Prentice Hall.

Dobrin. M.B., 1976. Introduction to geophysical prospecting. New York: McGraw-Hill.

Einsele, G. et al., eds., 1991. Cycles and events in stratigraphy. Heidelberg: Springer-Verlag, Inc.

Emiliani, C., 1992, Planet Earth-cosmology, geology, and the evolution of life and environment,.
New York: Cambridge University Press.

Gignoux, M., 1960. Geologie stratigraphique, Paris: Mason & Cie.

Howell, J.A. and Aitken, J.F. (editors), 1996. High Resolution Sequence Stratigraphy: Innovations
and Applications, Geological Society Special Publication 104

*Peyton, C. E., ed., 1977. Seismic stratigraphy-applications to hydrocarbon exploration. AAPG


Memoir 26. Tulsa, OK: AAPG.

SEPM, 1995. Special Publicaiton no. 53.

Schlee, J. S., ed., 1984. Interregional unconformities and hydrocarbon accumulation. AAPG
Memoir 36. Tulsa. OK: AAPG.

Sequence stratigraphy and facies associations, 1993. international Association of


Sedimentologists, Special Publication 18.

*Sequence stratigraphy as an exploration tool; Concepts and Practices in the Gulf Coast, 1990.
Society of Economic Paleontologists and Mineralogists Foundation, Eleventh Annual Research
Conference Gulf Coast Section, Houston, Texas.

*Van Wagoner; J.C., R.M. Mitchum, K.M. Campion, and V.D. Rahmanian, 1990. Siliciclastic
sequence stratigraphy in well logs, cores, and outcrops. AAPG Methods in Exploration Series
Number 7. Tulsa, OK: AAPG.

Weimer, P. and M.L. Link, eds., 1991. Seismic facies and sedimentary processes of submarine
fans and turbidite systems. New York: Springer-Verlag, Inc.

*Weimer; P. and H.W. Posamentier eds, 1993. Siliciclastic sequence stratigraphy: recent
developments and applications. AAPG Memoir 58, Tulsa, OK: AAPG.

*Wilgus et al., eds., 1988. Sea-level changes: an integrated approach. Society of Economic
Paleontologists and Mineralogist Special Publication 42. Houston, Texas. Note: this publication
includes a fold-out cycle chart with references.

Williams, G.D. and A. Dobb. eds., 1993. Tectonics and seismic sequence stratigraphy, 1993. The
Geological Society. (Also available from AAPG.)

* Texts marked with an asterisk are of special interest.


III. SANDSTONE RESERVOIRS

1. NON MARINE SANDSTONE RESERVOIRS

1.1 Classification

General Considerations

Figure 1 outlines a classification of the major continental environments in which sand is being
actively deposited. Figure 2 offers a generalized illustration of these environments. The
first three of these are mainly deposited by streams and can be placed in an order that roughly
corresponds to increasing distance from an upland sediment source area. They represent
environments in which progressively finer, more water-worked sediment is moved and deposited.
In contrast, eolian dune sands exist in a unique, nonaqueous setting, while lacustrine
environments often combine sedimentation regimes of many types, from deltas to submarine
fans. Alluvial fans and braided streams are generally the higher energy aqueous environments,
where lateral and vertical changes in lithology can occur suddenly and often. A number of
interrelated factors determine whether channels meander or braid. In general, braided rivers form
where gradient is relatively steep, discharge is occasional and heavy, and sediment supply is
high. Meandering rivers are more characteristic where gradients are lower, discharge more
perennial, and sediment is more fine-grained, with a higher proportion of silt and clay.

Nonmarine vs. Marine: Rules of Thumb and Their Limitations


Usually, there are no quick and easy methods that distinguish nonmarine from marine rocks. A
few general rules of thumb should be mentioned, however.In the sedimentary record, many non-
marine rocks are associated closely with unconformities, often lying immediately over, or being
sandwiched between, major erosional surfaces. Sediments deposited above the water table have
their pores filled with air and intermittently percolating groundwater, which is acidic and oxidizing.
( Figure 1 , Sketch showing relationship between local water table and oxidation. Iron-bearing
sediments within the phreatic zone have a much higher chance of taking on red coloration.)

They are, therefore, often orange or reddish in color, due to the presence of oxidized (ferric)
iron. Below the water table, however, pores are saturated with water that is often stagnant and
reducing. Consequently, iron tends to occur as a sulfide (e.g., pyrite) or, more commonly, in the
drab gray-green ferrous state. Thus, upland eolian, alluvial fan, and braided stream deposits are
often reddish in color, while lowland flood plain and lacustrine sediments more closely resemble
the drab color of most marine rocks.

Due to lateral lithologic variations in many nonmarine facies, seismic profiles through these rocks
are sometimes characterized by discontinuous reflections. ( Figure 2 , Seismic section to illustrate
the discontinuous reflectors which often characterize nonmarine deposits. Part a is a close-up
of the type of reflections commonly seen. Part b is a seismic profile of a preserved river channel.
Note discontinuity of reflections within the channel.) This, however, is a generalization and often
not the case for nonmarine sequences that represent considerable portions of geologic time.
Moreover, lacustrine sediments provide a conspicuous and important exception.

Other potential indicators are nonmarine fossils (e.g., vertebrates, freshwater mollusks and
gastropods, insect tracts, rootlet horizons) and exposed-surface sedimentary structures such as
raindrop imprints and mudcracks. Nearly all these may occur in transitional marine (e.g., deltaic)
environments as well, how-ever. In addition, these types of evidence are all distinctly rare in
outcrops, and therefore extremely so in cores.

As a whole, therefore, even in combination, these nonmarine indicators are suspect. In fact, none
of them may be present in a section which is determined by detailed analysis to be terrestrial. At
the same time, their presence can be used as guiding evidence to tip-off the geologist to the
possibility of a nonmarine facies.
1.2 Alluvial Fan Environments

Introduction
Alluvial fans represent an environment of sudden transition, where narrow, steep mountain or
upland valleys disgorge sediment onto an alluvial plain. Rapid erosion supplies coarse, often
bouldery detritus, which builds a debris cone out from the mountain base. ( Figure 1 , Block
diagram showing alluvial fan deposition in a semiarid to arid environment.) Thus, both on the
earth's surface and in the sedimentary record, fans are commonly banked against structurally
complex mountain fronts.

At times, they are juxtaposed with older, highly deformed sediments; at times, directly with
basement. In gross compositional aspect, fans will usually show an upward progression that
reveals the order in which material of the rising mountain block was attacked by erosion.

Alluvial fans are often referred to as syntectonic ("syn" = during) deposits, being generated by
tectonic regimes of rapid uplift, and should be expected to have developed in almost any setting
where orogenic activity has occurred inland from coastal areas. Where aggressive mountain
building has taken place along the margins of marine basins, fans commonly prograde from a
strictly terrestrial setting into a more transitional-marine regime, known as a fan delta. In arid and
semiarid climatic settings, alluvial fans often develop without significant depositional interruption
until they coalesce laterally to form bajadas. These have extensive, sheet-like geometries with
internal depositional complexity. Fanglomerate is the rock term employed for the coarsest, and in
some ways, most diagnostic, proximal lithology of the fan environment.

Much of the sedimentation that takes place on the fan itself is done by the work of ephemeral
braided streams; accurate analysis, therefore, recognizes a normal, though partial,
correspondence between the alluvial fan and braided alluvial environments. For the sake of
relative convenience, we treat these facies separately. The student should, however, be aware
that this introduces a certain degree of artificiality into normal facies analysis.

In the sedimentary record, preserved fan deposits are particularly evident in deposits of Late
Paleozoic — Early Mesozoic and Late Mesozoic — Mid-Tertiary age. These represent times of
significant tectonic activity throughout much of the world. In North America, the Pennsylvanian-
Permian time span is especially characterized by fan and fan delta development in the
Appalachian, Rocky Mountain, and New Mexico-Texas-Oklahoma regions.

Generally speaking, the style of deposition on alluvial fans prevents them from acting as good
reservoirs. To date, there are only a few, clear-cut examples of fields producing from terrestrial
fan facies. (Fan deltas, meanwhile, are productive in many parts of north Texas and southern
Oklahoma. Refer to Moore (1979) for an overview of Late Paleozoic paleogeography in this
region.) They are often, however, extremely important to recognize and delineate in the
subsurface due to their indication of both tectonic setting and source area composition. Given
this, and the fact that fans commonly grade into alluvial plain environments — whose sediments
have far greater potential to be good reservoirs — alluvial fans can serve as associated fades of
crucial significance.

Depositional Processes and Lithologic Characteristics


Sedimentation on an alluvial fan takes place mainly by a combination of:

• braided stream channel flow

• debris and mud flow (gravity sliding)

• sheet flooding

Most modern and ancient fans, if adequately mapped, can be divided into proximal and distal
subfacies. For the first of these, proximity to the sediment source, very steep drainage slopes,
and often irregular, but high discharge makes for deposition of detritus that is typically an
unsorted mixture of boulders, gravel, sand, clay, and possibly organic (soil) material. ( Figure 1 ,
Plan view of cross section of a single canyon and the alluvial fan which developed at its mouth,
Van Horn Sandstone, Precambrian (?) Texas. Shown are the downfan decrease in slope and
grain size, as well as the higher concentration of sedimentary structures in the distal portions. The
overall progression grades from massive conglomerate/gravel of the proximal fan, through
alternating trough and foreset crossbedded sand of the midfan, to finer-grained, more thinly
bedded crossbedded sands of the distal fan.) This type of sedimentation occurs right up against
the mountain front itself. In addition to high energy stream deposition, landsliding, in the form of
debris and mudflow, is also common. Down-fan, where braided channel flow begins, more
gravelly beds alternate with crossbedded sands. With greater distance from the mountain front,
the proportion of sand increases, and the shallow cut and fill of braided channel deposits
predominates. Mudflows may occasionally spread finer-grained clay material over parts of the
fan.

Preserved alluvial fan complexes, therefore, usually show the overall lateral and vertical
progression shown in Figure 1 . Vertical sections are usually dominated by a large number of
stacked sequences that each show an upward change from conglomerate to coarse and medium-
grained sand. Each sequence is truncated abruptly by the next overlying sequence, indicating the
lateral shifting of the braided channels over the fan. Most identification of fan facies in the rock
record, however, has concentrated on the distinctive fanglomeratic lithologies. ( Figure 2 ,
Photograph of fanglomerate, Sawatch Mountains, Colorado. This particular section was part of a
fan system that spread from the basement block uplifts of the Ancestral Rocky Mountains in
Pennsylvanian time.)

A number of specific sedimentary structures will be found in fan sediments. These will be
concentrated in channels, where current action is most dominant. Cross-bedding, pebble
imbrication and long axis orientation, current lineations (from smaller grain orientation, for
example), dune and ripple morphology are those most commonly found. In addition, the long
dimension of channels can be used to derive paleocurrent patterns. These will show the overall
morphology of the fan, since most stream flow is directly down the steep slopes, i.e., "fanned"
out. Detailed paleocurrent information is not often systematically collected for fan deposits. The
downfan (basinward) direction is usually fully evident from large-scale stratigraphic parameters,
or from seismic and structural data.

Paleontology
As might be surmised, macrofossils are relatively rare in alluvial fan complexes. Some vertebrate
bones and plant fragments may exist, the result of sudden burial due to flooding. Normally,
however, the depositional caprice of the fan environment prevents any significant floral or faunal
colonization. Microfossils, such as spores and pollen, have a far better chance of being
incorporated. At the same time, percolating groundwater within the fan would be acidic and these
often-valuable remains have a strong chance of being strongly oxidized, thus partially dissolved
and often altered beyond recognition.

Geometry
The size and detailed morphology of fans will vary according to climate, source rock, basin
architecture, and the degree and style of tectonic activity. Individual fans are normally wedge-
shaped in plain view and concave-upward in cross section. ( Figure 1 Plan view of cross section
of a single canyon and the alluvial fan which developed at its mouth, Van Horn Sandstone,
Precambrian (?) Texas. Shown are the downfan decrease in slope and grain size, as well as
the higher concentration of sedimentary structures in the distal portions. The overall progression
grades from massive conglomerate/gravel of the proximal fan, through alternating trough and
foreset crossbedded sand of the midfan, to finer-grained, more thinly bedded crossbedded sands
of the distal fan.) They may also reveal a lens-like morphology in the subsurface. Fans can have
radii that range from hundreds of feet to tens of miles. Bajadas can be over 100 miles long,
though considerably narrow in width. Sections have been measured up to several thousand feet
thick, with separate conglomerate-sand sequences showing lateral continuity of up to about 1500
ft. Clay layers are much more localized and thin. Commonly, a fan or fan complex will be thickest
at its proximal end, having, for example, filled in a rapidly subsiding local graben.

Associated Facies

Since many fans grade laterally into alluvial plains ( Figure 1 , Distribution of different rock
types in an actual alluvial fan complex of eastern Pennsylvania and Figure 2 , Cross sections of
alluvial fans showing common associated facies. Note that interfingering with braided stream
alluvium may be subtle, barely detectable), their more distal channel deposits and sheet-flood
material are not easily distinguished from those of braided streams. In general, alluvial deposits
are characterized by sediments of finer-grain size, better sorting and rounding, and more
abundant sedimentary structures. In arid environments, fans may be associated with the local salt
deposits of ephemeral playa lakes. As discussed, the proximal end of most fans is directly
adjacent to a mountain front, and this conspicuous association is usually considered to be
diagnostic.
Diagnostic Evidence

Seismic

Vertical seismic sections through alluvial fan complexes typically show discontinuous internal
reflectors. ( Figure 1 , Seismic section through a probable alluvial fan that developed over
structurally deformed basement. Note the relatively poor internal seismic character to the
deposit.) This should be expected, given the great lateral and vertical variation in lithology.
Furthermore, the concave-upward ideal profile of an individual fan is more often subdued in the
subsurface, due to postdepositional compaction and tectonic tilting. Bajadas are more extensive
and sheet-like and will display only a vague fan-type profile over their width. Where fan
development continued for considerable periods of time, the original mountain front may be worn
down and completely buried in its own debris; in this case, fan sediments are likely to grade
upward into braided stream alluvium, which is also characterized by poor internal reflections.
Seismic profiles can be very useful in delineating old mountain fronts and often the extent of the
immediately adjacent sediments they produced. Specific lithologic types within the fan itself,
however, are not normally discernible.

Wireline Logs

Most geologists maintain that alluvial fan facies generate no important diagnostic response on
wireline logs. The diverse processes acting to create and prograde the fan result in considerable
vertical and lateral change, which means that log profiles will be correspondingly varied and
therefore difficult to correlate. On the other hand, it should not be overlooked that this type of
unpredictable log response showing substantial changes between even closely spaced wells,
might itself be considered a characteristic of fan deposits.
Dipmeter logs ( Figure 2 , Idealized gamma ray and dipmeter logs for an alluvial fan sequence,
showing both fanglomerate and channel development. Note the three major patterns: lowest
green dips, which represent shale breaks and correspond to spikes on the gamma ray, curve;
random "bag o'nails" dips in fanglomerate; and dip clusters that show an upward-increasing blue
pattern in channel sands.) will most often show a "bag o' nails" pattern for fanglomerate. A green
("shale", i.e., structural slope) pattern may be discernible, but depositional slopes are frequently
high (5° to 25° ), and this could confuse correlations. Due to the braided stream flow on a surface
of radiating geometry, dip switches are characteristic. Frequency diagram plots, ideally, should
show approximately a 180° azimuth sweep with, perhaps, a subdued bisecting maximum that
represents the influence of the feeding stream.

Figure 2 shows an idealized dipmeter profile through a fan. Several shale layers are indicated
simultaneously by spikes of high radioactivity on the gamma ray log and low dips (green motif) on
the dipmeter log. These shales separate three channels, whose tadpole patterns show a
clustering at a higher magnitude of dip. This is due to cross-bedding. These channels overlie the
random "bag o'nails" pattern of fanglomerate, and thus show the outward progradation of the fan.

Cores

The general coarseness, poor-sorting, clast-angularity, and immaturity of alluvial fan sediments
are the conspicuous features which dominate most core samples of the upperfan. Finer-grained,
cross-stratified or flat-bedded channel sandstones can also be prevalent, particularly from midfan
sediments. It is therefore, the association of these — as well as that of thin shales (mudflows),
sand-silt beds (sheet-floods), and gravel layers — which is most prescriptive in terms of what is
normally seen in core samples. Such diversity might be found in cores taken from a single well, or
from a number of nearby wells.
1.3 Braided Stream Environments

General Introduction

As we mentioned earlier, many present-day alluvial fans grade laterally into the alluvial plain of a
braided river; that is, one characterized by an interlacing, vein-like network of low-sinuosity
channels with constantly shifting midchannel bars. ( Figure 1 , Block diagram model of a
braided stream system in a semiarid environment and Figure 2 , Shifting channels in a braided
river course -Durance River- near Avignon, southern France, between 1939 and 1958. )
Streams and rivers tend to braid when three main factors conspire: (1) high (though possibly
seasonal) discharge, (2) relatively steep slopes, (3) large amounts of coarse sediment.
This style of deposition is imposed basically because more coarse sediment exists within the
environment than the river or stream can carry at any one time. As shown by Figure 3 (Plan view
of cross section of a single canyon and the alluvial fan which developed at its mouth, Van Horn
Sandstone, Precambrian (?) Texas. Shown are the downfan decrease in slope and grain size,
as well as the higher concentration of sedimentary structures in the distal portions. The overall
progression grades from massive conglomerate/gravel of the proximal fan, through alternating
trough and foreset crossbedded sand of the midfan, to finer-grained, more thinly bedded
crossbedded sands of the distal fan.), braided streams often begin on the fan itself. Full-sized
braided rivers occur in a variety of specific settings, including broad semiarid plains, valleys
longitudinal to mountain fronts, and in glacial environments, for example at the edges of ice caps.

In the rock record, braided channel systems grade into fanglomerates at their proximal end.
Distally, they may terminate in desert lakes, sand dune complexes, meander channel flood-
plains, or fan deltas building into lakes or seas.

Depositional Processes and Lithologic Characteristics


The overall discharge of a braided stream is low compared to the supply of sediment. It is also
sporadic — punctuated by extreme fluctuations that range from no water at all to torrential
flooding. Slopes are relatively high (usually 3° to 10°) and most sediment is coarse-to medium-
grained sand, with sometimes major amounts of gravel and more minor contributions of silt and
clay. The preservation of this environment requires that the braiding river flow through a region of
subsidence, for example a developing basin, in relative proximity to a rising orogenic front. This
condition exists in a number of specific tectonic settings, such as intermontane, subsiding coastal
and rift graben basins. Like alluvial fan deposits — with which they are often genetically
connected — braided stream complexes are especially common in sediments of Late Paleozoic
and Middle Mesozoic-Early Tertiary age, when tectonic activity was extensive over large regions
of arid or semiarid climate.
Modern braided stream deposits are well studied. The shape of individual bars can be either
tabular and linear, or curving. However, complete morphology is rarely preserved in the
sedimentary record; most bar and channel sequences show a great deal of internal truncation. (
Figure 1 , Surface features and facies model of a braided channel system. Sedimentation
takes place almost entirely within the shifting complex of channels with some small silt deposition
in abandoned channels. No flood plain exists; migration of the channel complex, due to tectonic
tilting, for example, leads to the creation of extensive sand bodies.)
Floodwaters will quickly overflow the shallow channels and cover the entire alluvial plain. During
this time, new channels will be scoured and cut, and old ones filled. Bar migration occurs as
discharge increases and then decreases: upstream portions lose sediment to rising currents,
while the downstream end and flanks build as the water level and carrying capacity of the stream
diminish. A vertical section through a braided alluvial complex, then, will show a high degree of
cut and fill. ( Figure 2 , Photo showing cut and fill channeling and general character of a recent
braided stream deposit.)

Some deposition of finer-grained material can also occur, particularly in channels abandoned
for longer periods of time. In more humid regions, vegetation may colonize some bars and
stabilize them with soil cover. Silt and clay that settle in abandoned channels are commonly
ripped up during flooding and may be preserved as intraformational conglomerate.

Vertical sequences are commonly comprised of multiple, stacked channels, each showing an
overall progression that begins with the sharp erosional base of a channel floor ( Figure 1 ).
Above this, in ascending order, are gravelly, conglomeratic debris with imbricated, subrounded
pebbles; coarse, crudely bedded sands with some gravel; and medium-grained sandstone that
fines upward. The more coarse, pebbly sands often compose as much as 95% of the total
lithology. Those sandstones without gravel show lamination, planar-bedding, cross-stratification
of both trough and tabular types, and where they are more fine-grained, small-scale ripple
development. Due to rapid changes in water level (which causes sudden changes in intergranular
pore pressures) soft sediment (also called "quicksand") deformation is also common in braided
alluvial deposits and results in convoluted bedding from both water-escape and recumbent or
collapsed cross-bedding. Sorting ranges from poor, near the base of such a sequence, to very
good at the top. However, the continual cutting and filling of channels ensures that very few
complete sequences exist and, therefore, that sudden lateral and vertical variations in sorting and
grain size are characteristic.
The lack of fine r-g rained material theoretically means that porosities and permeabilities should
be high; however, the above-mentioned changes in sorting and size, as well as diagenetic
influences, often reduce these parameters considerably. Figure 3 (Plots of porosity versus
permeability showing rapid vertical change, Elk City field, Oklahoma. ) gives two examples of
porosity-permeability profiles through a rather narrow producing interval that shows substantial
and unpredictable variation. Such profiles are extremely useful in indicating local, high-
porosity/permeability zones.

Paleontology

As with alluvial fan lithologies, braided stream deposits are often colored red, orange, or yellowish
by varying amounts of ferric iron. A second similarity is their general hostility to plant and animal
life. Sudden flooding, which inundates and changes the specific configuration of channels and
bars, prevents most colonization and also flushes out what organic remains may have settled on
the immediate alluvial plain. Furthermore, groundwater percolating through buried sands is
oxidizing and acidic, and might be expected to destroy or disfigure much original paleontologic
evidence. Some burrowing or rootlet horizons may be preserved in shale layers, but on the
whole, these sands are commonly barren and difficult to date.

Geometry

The geometry of braided stream deposits is usually sheet-like and can be extensive where
numerous rivers were active or if tectonic tilting of the regional depositional surface caused the
original river to migrate laterally. Margins may be irregular due to regional stratigraphic changes
such as pinch out. This type of deposit will often be long and rather narrow, however. As shown
by deposits at Prudhoe Bay and in the Sirte basin of northern Libya, total thickness can vary from
hundreds to thousands of feet. Sandstone-conglomerate sequences enclose thin, discontinuous,
relatively rare shale lenses. ( Figure 1 , Surface features and facies model of a braided channel
system. Sedimentation takes place almost entirely within the shifting complex of channels with
some small silt deposition in abandoned channels. No flood plain exists; migration of the channel
complex, due to tectonic tilting, for example, leads to the creation of extensive sand bodies.)
Associated Facies

As mentioned, alluvial fans form a common associated facies at the proximal end of braided
stream deposits. At its distal termination, the modern braided river grades into a diversity of
environments, including meandering alluvial plains, and various arid or semiarid coastal
depositional regimes. In the case of the latter, association with evaporitic facies — either
ephemeral playa lakes, eolian dunes, or the dry coastal environment of the wadi-sabkha, the best
developed today in parts of northern Africa along the Persian Gulf.
( Figure 1 , Schematic cross section through the Trucial Coast of Oman, showing transition from
alluvial fan to coastal marine environments. Braided stream/wadi facies would occur between
the fan and dune facies.) The wadi fluvial plain, largely braided, is determined by a short period of
seasonal discharge that creates short-lived streams and extreme flooding, which subsequently
dries up completely. The wadi passes into the highly evaporitic, transitional marine sabkha, which
largely consists of dolomitized carbonates, algal mats, and gypsum/anhydrite mush deposited in
an arid salt marsh (Selley 1978).

Diagnostic Evidence

Seismic

Due to relative lithologic homogeneity, braided stream deposits do not often show internal
reflections. Shales are too thin and localized to generate any significant responses. Figure 1
(Possible braided stream/alluvial fan deposit in seismic section) is a probable example of the
overall lense-like geometry and "poor" internal seismic character of such a deposit. At the
same time, if the facies is thick enough — that is, if it represents sedimentation for much or all of
a geologic period, certain large-scale depositional changes will most likely have taken place, and
thus the interfaces these have created will generate continuous reflections. This is shown by a
seismic section of Prudhoe Bay field, where an apparent braided river facies composes the entire
Permo-Triassic section. ( Figure 2 , North-south seismic section through the discovery well at
Prudhoe Bay. Note the detail of interpretation possible on this section. Updip truncation of the
Permo-Triassic section against the unconformity is clearly seen.)
Wireline Logs
Figure 3 (Log of a braided alluvial sequence showing characteristic monotonous log response.
Note that the gamma log is neither as clean as that for eolian deposits, nor as shaley as
meander channel flood plain alluvial sequences. Azimuth frequency plots reflect the linear trend
of this type of river.) displays the somewhat idealized log response of a braided stream deposit.
Some crude fining-upward portions of the curve can be discerned, but grain-size variation is most
often too small to produce a convincing bell-shaped channel pattern. Instead, the gamma ray
curve normally remains monotonously at lower values, with occasional spikes. This can usually
be interpreted as a section composed mainly of coarse sediment, encasing a few thin, finer-
grained interbeds. Though the mostly "massive" profile is not singular to braided stream deposits,
it does indicate an overall sedimentary character that should clue in the exploration geologist to
this possibility.

In terms of its dipmeter signature, this facies mainly shows the multiple stacking of channels.
Within each channel, azimuths and dip amounts are clustered into separable groupings. Channel
switching is, of course, characteristic, but azimuth changes will usually remain within a 90° arc,
since individual channels are not highly sinuous and will not wander far from the overall
downstream direction. Consequently, this direction — and the probable long dimension of the
sand body as a whole — can often be found by bisecting the arc when it is plotted on an azimuth
frequency diagram. If little tectonic disturbance has occurred, dip amounts will vary from near
zero in planar beds to above 40° in the case of convoluted bedding. Most crossbed sets will
range from 5° to 25°, with tabular types representing the higher end of this range.

Cores

Core samples taken from a braided alluvial section can reveal either a homogeneous section of
coarse, crossbedded and gravelly sandstones, or a diverse range of grain sizes and sedimentary
structures. Again, well-preserved individual sequences begin with a sharp erosional base that
marks the channel floor and is overlain by an upwardly-fining progression of grain sizes and
sedimentary structures. Figure 4 (Idealized "outcrop" showing succession of grain sizes and
sedimentary structures in a single channel sequence of a braided alluvial system) shows the
general relationships between grain size and sedimentary structure.
Basic Reservoir Characteristics
The alluvial deposits of braided rivers occur as blankets of coarse, sandy sediment. Grain
rounding is moderate at best, and sorting is often intermediate to poor; thus, despite high, gross
porosity, these deposits typically show a good deal of permeability variation. At the same time,
due to sediment coarseness, intercommunication between individual sandstone bodies (channels
and bars) is usually excellent. Due to their continental setting, braided alluvial deposits are
normally capped by an unconformity in the sedimentary record, and as a result, often occupy
stratigraphic positions with good reservoir potential.

Postburial percolation of acid meteoric waters through this type of deposit has almost an equal
chance of destroying primary pore space by mineral precipitation as enhancing it through
dissolution. In cases where such fluids have been especially active, (usually just below an
erosional surface), an upper "leached zone — within which cemented material has been partly
removed — may overlie a tight interval where reprecipitation of this material has occurred. A high
degree of "tightening" due to this type of early diagenetic process is shown in northern Libya's
Sirte basin, where Paleozoic and Cretaceous non-marine sandstones are much more silicified on
structurally elevated basin flanks and highs (in areas closer to the major unconformity) than in the
deeper, basinal region. We'll look more closely at this particular setting below; it forms one of the
several giant field areas that produce from braided alluvial sands.

As discussed, shale layers in this facies are usually inconsequential in terms of fluid migration
blockage, and therefore, do not promote or aid stratigraphic trapping. Neither do they serve as
good source rocks. Moreover, the development of braided stream systems relatively far inland,
deep within continental shield and intracratonic basin regions, limits their eventual accessibility to
petroleum generated from marine source rocks. The ideal reservoir setting for this facies is
immediately beneath a major unconformity over which rapid marine transgression has occurred. If
subsequent tectonism were to auspiciously juxtapose these potential host rocks against overlying
marine shales in a regime of high heat flow, so much the better. ( Figure 1 , Idealized cross
section to show how braided alluvial sand blankets may act as sites for large oil accumulations on
truncated horsts, sourced and sealed by transgressive marine shales. In such a setting,
tectonism-possibly rift-related-would generate the high heat flow necessary to mature organic
material into hydrocarbons.) In fact, in a very basic way, this has been the case at Sirte and at
Prudhoe Bay, another giant field area.
Examples
The two field area examples discussed below represent giant accumulations of petroleum
(greater than 1 billion bbl), and therefore show the impressive importance this facies has had to
oil exploration and development.

Sarir and Messla Fields, Sirte Basin of Northern Libya

The Sarir and Messla fields lie on the southeastern edge of the Sirte embayment of Libya. (
Figure 1 , Location of Sirte basin and its major oil fields.) Both are giant fields, with oil reserves of
3 and 12 billion bbl in place respectively (Gillespie and Sanford 1967; Sanford 1970; and, Clifford
et al. 1980). In both, the reservoir is the lower Cretaceous Sarir sandstone, which includes a
variety of nonmarine facies, but is dominated by braided alluvial deposits. It contains several units
that are truncated by an unconformity and onlapped by an Upper Cretaceous group of shales,
marls, and limestones (Rakb Group), which have acted as both the source and, to some degree,
the seal of the petroleum accumulations.

Porosity in both fields is not primary, but the result of fracturing induced by rift tectonism, which
created the basin. There is an average of two fractures per foot, as determined from core
sampling, and porosities average about 18% to 20%. Permeabilities average several hundred
millidarcies, but intervals of as much as 2 to 3 darcies occur. The main factor contributing to the
giant size of these fields, however, is the thick and continuous nature of the braided alluvial
reservoir sands.

The Messla field is an unconformity trap in which the Sarir sands are truncated against
Precambrian basement. ( Figure 2 , Cross section of Messla field with sample log of the
productive Sarir Group braided alluvial deposit.) It has a gross pay thickness of 90 ft and is about
375 km2 in aerial extent.
The Sarir field is more complex, being a combination trap determined by both the mentioned
unconformity and a block-faulted anticlinal structure.
( Figure 3 , Cross section of Sarir field, showing truncation of braided alluvium and other
continental deposits by a major unconformity. ) In size, the producing area is somewhat larger
(about 420 km2) than the Messla field, and its oil column is considerably thicker (300 ft).
Prudhoe Bay

The well-known Prudhoe Bay field of Alaska's North Slope is another example of a giant field
whose major reservoir is a braided river sand capped by an unconformity and transgressive
marine sediments. Though somewhat similar, the specific setting is more complex than at Messla
and Sarir. A thick section of Late Paleozoic to Jurassic-age deposits were deformed and uplifted
along a rising basement high in Late Cretaceous time.
Subsequent marine incursion buried the erosional surface, against which were truncated all older
formations, including the Permo-Triassic Sadlerochit sandstone, a coarse, pebbly, crossbedded,
braided alluvial deposit. Trapping in the Sadlerochit is, as at Sarir, due to truncation at the updip
truncation on an anticlinal limb beneath the unconformity and against a normal fault. Figure 4 (SP
and resistivity log of the Permo-Triassic reservoir interval in a productive Prudhoe Bay well) gives
the log and lithologic profile seen in a producing well. Figure 5 (Structure contour map of the
producing sandstone at Prudhoe Bay. Star symbols indicate successful wells in the Permo-
Triassic. Faulting is simplified; contours are in feet. ) reveals the general complexity of the field
Prediscovery seismic data was of excellent quality. ( Figure 6 , North-south seismic section
through the discovery well at Prudhoe Bay. Note the detail of interpretation possible on this
section. Updip truncation of the Permo-Triassic section against the unconformity is clearly seen.)
Clearly seen were the updip wedge-out of the potential reservoir formation, as well as its specific
structural setting. As a consequence, the discovery well was ideally located right at the crest of
the Prudhoe Bay structure, and penetrated more than 400 ft of gas cap before encountering oil.

Reservoir properties are truly excellent, with 20% to 25% porosity and permeabilities ranging from
300 md up to 3 darcies. Estimated original recoverable reserves were 9.6 billion bbl of oil and 26
tcf of natural gas.
1.4 Meandering Channel Alluvial System

General Introduction
With greater distance from the sediment source area, a meandering river becomes more typical. (
Figure 1 , Idealized block diagram showing a meandering river system over a region of low slope
and continual subsidence. ) Alluvial flood plains cut by a single meander channel occur in
regions characterized by relatively low gradients, fine- to medium-grained sediment, and
continuous (nonseasonal) discharge. The boundary between braided channels and a single
meander channel is a gradational one, but the distinction is very important for petroleum geology.
We have seen that braided systems deposit extensive blankets of sand, some which have served
as reservoirs for giant accumulations. Meander channels, by contrast, deposit more localized
reservoir sands, which host generally small, elusive stratigraphic traps. Such rivers very often
create deltas at their mouths, and thus much of the following discussion on deposition, geometry,
and identification has relevance to channels in this more transitional environment. In the
sedimentary record, a large number of preserved meander channel facies are associated with
deltas.

Depositional Processes and Lithologic Characteristics

Figure 1 illustrates the basic sedimentary model for meandering channels in an alluvial flood
plain. As shown, there are three main depositional subenvironments: the active channel, the
abandoned channel, and the flood plain itself. As with braided streams and alluvial fans, flooding
is the main process that causes deposition within each.
The creation of sand bodies occurs as a result of two factors: channel migration, with the
simultaneous deposition of point bar sand; and, channel abandonment, where point-bar sands
are buried beneath finer-grained, backwater material in oxbow lakes. The former is by far the
most important to petroleum geology, since the latter results in very thin, isolated sand layers.

Meandering stream channels display an asymmetrical profile which reflects the distribution of
current velocity and thus deposition and erosion ( Figure 1 ). Basically, flooding causes erosion of
outside meander ("cut") banks as current strength rises, and subsequent deposition on the inner
banks as floodwaters recede. Point bars build out into the channel, and their progradation mostly
keeps pace with the erosion of the opposite cut bank. The specific direction of their accretion is
largely at right angles to the downstream current at any one point. Point bars are characterized by
a fining-upward sequence of grain sizes and sedimentary structures that reflect the decreasing
current velocity from the gravelly channel floor to the silts and muds of the flood plain. ( Figure 2 ,
Idealized "outcrop" showing upward succession of grain size and sedimentary structures in a
preserved point bar.) Most have slopes of 10° to 20°. They are the site where the majority of
sand deposition fakes place in the alluvial systems of meandering streams.

An abandoned channel results when a meander is cut off. This is, to some extent, the inevitable
culmination of meander development. Continual cut bank erosion both increases the amplitude of
individual loops and "tightens" their necks. ( Figure 3 , Sequential diagram to show how oxbow
lakes, i.e., abandoned meanders, form as a result of continual meander migration. Erosion of
the outside cut bank occurs simultaneously with deposition of sand on the point bar of the inside
bank. Abandoned meanders are eventually filled in with a plug of clay and silt. Most meander
belts, therefore, are composed of both shoestring point-bar and oxbow deposits.) At some point,
flooding will cause currents to overflow the point bar and erode a new, straighter channel across
it, thus decapitating the old meander bend. Thereafter, the sinuous oxbow lake which remains will
be filled in by finer grained material, mostly introduced by subsequent floodwaters. The point-bar
sequence will thus be incomplete, its lower, coarser sand and gravel overlain by silts and clays (
Figure 1 ).
The flood plain itself is composed predominantly of backwater muds (possibly including swamp
peat), which are intermittently blanketed by thin layers of fine sand and silt deposited by periods
of overbank discharge. Along the banks of the river are two other, localized floodplain subfacies:
levees and crevasse splay deposits. Levees form as a result of floodwaters dropping their
heavier, coarser load when they first overflow the channel and enter the flood plain. Crevasse
splays occur when these same waters break through preexisting levees to spread a cone of
sandier material further out from the river banks.

Channel migration is largely unpredictable in its specific evolution, and thus the location and
vertical arrangement of point-bar sands within a preserved stratigraphic section is often difficult to
determine. Furthermore, as with braided stream deposits, full sequences for individual channels
are not always seen in the record. Continual lateral migration of channels from one side of the
flood plain to the other means that point-bar deposits will suffer a certain degree of cut and fill.
Usually, however, this is far less pronounced than in the case of braided stream alluvium.
Individual point-bar or abandoned channel sands form relatively isolated "shoestring" bodies,
encased within impermeable flood-plain deposits. ( Figure 4 , Isopach map and east-west cross
section of two stream channels in the Muddy Sandstone of South Glenrock oil field, Wyoming-
southwestern Powder River basin. Note how two channels fill a single meander belt. Oil is
stratigraphically trapped in point-bar sandstones within the channel fill.)
These subenvironments all exist within a single meander belt, which may range from tens to
thousands of meters in width. In the case of a subsiding basin, where the flood plain is broad yet
of low slope, tectonic tilting may force the river to either change course altogether ( Figure 5 ,
Idealized block diagram showing a meandering river system over a region of low slope and
continual subsidence), or to migrate laterally across the entire basin. In each case, new
bodies of sand will be created by the new meander belts, each composed of stacked channel and
point-bar sequences. Thus, meandering alluvial systems do not always result in shoestring
reservoirs.

Paleontology
Fossil remains most often preserved by this environment include plant debris, vertebrate bones,
freshwater mollusks and gastropods, and, in greater abundance, spores and pollen. Low-grade
coals derived from backwater swamp peat deposits may also contain some indicative remains.
Rootlet horizons and burrowing, as well as surface sedimentary structures, such as raindrop
impressions and desiccation cracks, are seen in modern-day settings and in outcrops, but the
chances of discovering any of these in cores is, at best, slim.

Geometry
The size of an individual meander belt has already been mentioned. Given an open basin setting,
this — as well as the frequency of meander migration — will be determined by climate, slope, and
sediment supply. Vertical thicknesses for preserved meander channel facies vary from about 10 ft
to 1000 ft (3 m to 300 m). As a whole, a meander belt will trend perpendicular to a paleoshoreline,
but this will normally become clear only where a great and regional abundance of well data exists.
It may, however, be used as an overall predictive concept if other large-scale factors, such as
valley or basin trends, become known. Most modern-day meander belts are limited to about twice
the average meander amplitude, and this can also be used as a general guide for diagnosing at
least the lateral extent of a given reservoir. As mentioned, shoestring sand bodies are a common
result of this facies. These can be less than 2 m or more than 30 m thick, and either continuous or
sporadic.

Associated Facies
Meandering stream deposits very often grade basinward into deltaic and shoreline-shelf facies,
which, in turn, pass into distinctive deeper water sands and shales. Lacustrine sediments are
another common associated environment. Of these, the deltaic association is probably the most
important for petroleum geology. Normally, the transition from the meandering channel flood plain
will be gradational if a delta is actively prograding out into a basin, and sharp if it is inactive (Cant
1982), i.e., being reworked by marine transgression.

As mentioned, braided streams often form the proximal associated facies for this environment.
Other common associations are with large inland lakes and mountain streams. In contrast to
many braided stream deposits, those of the meandering channel facies are normally devoid of
oxidized iron coloring, being greet and drab. The much higher clay component, smaller grain size,
and textural characteristics, such as better sorting, generally provides a basis for roughly
distinguishing the two facies where they grade into each other.

Diagnostic Evidence

Seismic
In the subsurface, channels generally create abrupt changes in lithology. Their seismic "visibility"
should, therefore, be pronounced. At the same time, where the sharp erosional base and sides of
the average channel make for good velocity contrast, the upper part of the average channel
grades into flood-plain deposits and thus will not generate high-quality reflections. As a result, the
typical lens shape of most channels should be only relatively clear on high resolution seismic
lines, as shown in Figure 1 . (Seismic expression of a river-cut channel. Note the abrupt
termination of flat-lying reflections against the channel flanks and the change in seismic character
between these reflections and those within the channel. Note also the steeper slope of the right
flank of the channel, possibly indicating that this was the cut bank.) Diagnosing such a channel,
however, does not identify the environment it was deposited in. Other information is required to
determine whether the specific facies is a delta estuary, submarine fan, or meandering stream
complex.
Since meandering, like braiding, is a lateral feature, a horizontal time slice through a block of 3-D
seismic data may reveal the map view of an ancient stream. ( Figure 2 , Horizontal time slice
showing the meandering channel geometry and oxbow lakes of an alluvial flood plain.) This is a
convincing example of the advantages possible in doing this type of seismic surveying.
However, choosing the correct time interval of a section that will show such detail is neither
simple nor accidental. An abundance of both geologic and seismic data must already exist
concerning the formation to be examined. Furthermore, the velocity spectrum of shale overlaps
that of high porosity sand and, in Figure 2 for example, it cannot easily be surmised what
comprises the channel fill.

Wireline Logs

Figure 3 (Well log showing two upward-fining point-bar sand bodies) gives a suite of logs showing
the variations and relationships in alluvial flood-plain sediments. Two upward-fining point-bar
sequences are in evidence. Both are surrounded by overbank flood-plain shales. Note how the
gamma ray (SP) curve shows the abrupt change from shale to sand at the base of each channel,
as well as the fining-upward, bellshaped curve as point-bar sand grades into flood-plain shale at
the top of each channel sequence.

The dipmeter log for such a section will be a bit complex, but will show three main depositional
surfaces:

• structural slope (green motif)

• major accretion slopes (red motif)


• cross-bedding dips (blue motif)

In relatively undeformed areas, the green ("shale") motif is the lowest in dip amount and indicates the
regional or structural slope over which the broadest depositional processes occur. The upward decreasing
red motif represents, in this case, the larger-scale accretionary layers, which show flatter and flatter dips as
the channel becomes filled in. These surfaces, as shown, dip in toward the channel axis, usually at right
angles, and consequently point towards the thickest portion of the individual bar. They represent major
successive surfaces of the point bar itself. ( Figure 4 , Idealized dip log pattern showing the progressively
lower slope amount (red motif) characteristic of filled-in channels. Tadpoles shown correspond to dips of
major accretion surfaces, in this case, those of the point bar.)

Cross-bedding dips within these accretionary layers will not be systematically or consistently read
by the dipmeter tool when it is set on a broad step distance and correlation interval. If this interval
is made smaller, however, crossbed sets become discernible as a third, upwardly-increasing dip
trend known as the blue motif. ( Figure 5 , Idealized dip log showing both the filled-in red motif at
left and the upward-increasing blue motif, which indicates individual crossbed sets. Note that
the blue pattern at right depends upon a narrow dip correlation interval-usually less than ten feet-
so that both toeset and foreset dips can be recorded by the logging tool.) Crossbeds are a current
direction indicator and will thus trend usually at right angles to the accretion slope, which, as
mentioned, dips into the channel. An azimuth frequency diagram for the dip log of a channel
should ideally, show a bimodel pattern, indicating both the lateral point bar accretion and the
cross-bedding within it. The latter will show a higher frequency and reveal the principal
downstream direction — but only at a particular point. Due to the meandering nature of channels,
current direction will change as much 180° to 270° between locations. Therefore, only the most
limited prediction for reservoir alignment can be made on the basis of dipmeter data from
individual wells.
In general, log data can be used to construct isopach maps which, when combined with
paleocurrent information, are an often-used base for geologists in predicting field size and trend,
as well as the possible locations of upstream or downstream channels with similar reservoir
properties. Various units may be contoured. In addition to the reservoir sand itself, the underlying
unit into which the channel cuts is also a good choice for delineating reservoir geometry and
extent. It is sometimes useful as a first step to outline the general location of the valley or plain
into which the channel once cut. Isopaching the channel sand interval is very often one of the
best ways to articulate the original river course on maps. Overall, the geologist should keep in
mind that channels create sudden changes in subsurface formations. Such changes should be
detectable through a variety of techniques.

Cores

Core sampling of channels should show the range and sequence of sedimentary types and
structures mentioned. However, such sequences are often truncated by overlying channels and
the entire suite may not be seen.
If sufficient sampling is done, the geologist will be able to roughly predict reservoir continuity. One
rule of thumb is based on the general 1:1 sand-to-shale ratio of meandering alluvial facies that
reflects near-equal development of flood-plain and point-bar subfacies. This rule is shown
schematically in Figure 6 . (Two overall schemes depicting potential vertical continuity of
alluvial channel sand deposits. When the sand/shale ratio for an alluvial deposit is consistently
low-less than 1-, it may be assumed that flood-plain sediments dominate and point-bar sand
bodies are relatively isolated. Conversely, when this ratio exceeds unity and sands are coarse,
the probability is higher that channel sands are abundant and interconnected.) When shale (i.e.,
the flood plain) composes more than 50% of the vertical section, channels are more likely to be
unconnected; reservoirs, therefore, should be smaller and more numerous. When sand (i.e., the
point bar) dominates, channels have a higher chance of being stacked and interconnected. In
addition, the coarser the sand becomes in this latter case, the greater the probability for extensive
cut and fill, and therefore vertical intercommunication.

Reservoir Characteristics
Because of their high shale content and the often restricted size of individual point-bar deposits,
meander channel systems are seldom able to provide laterally extensive reservoirs capable of
hosting giant petroleum accumulations. Generally, they support smaller fields, many of which can
be found only by serendipity. Most trapping is either stratigraphic or a combination of stratigraphic
and structural influences. Lateral fluid flow is limited by:

• surrounding flood-plain shales (extensive)

• abandoned channel clay plugs (local)


Figure 1 (Isopach map and cross section of the Cretaceous Denkman and Little Creek field, Mississippi.
Note the scale of the meander loops and the lenticular nature of the sand body.) shows the lenticular
geometry of a typical field in point-bar sands.

A potential drawback that meandering channel complexes share with braided alluvial deposits is
that they are often deposited a considerable distance from marine source rocks. However, since
many flood plains pass laterally into coastal environments, their channels have sometimes acted
as the highest, updip traps for petroleum that has migrated out of marine basin depocenters. In
addition, flood plains are sometimes buried beneath transgressive marine sediments, which can
serve as both source and seal rock. Alluvial muds are themselves also capable of sourcing a
variety of gas-generating kerogen.

Examples
The vast, shallow interior seaway of the Later Cretaceous divided the North American continent
essentially in half, and encouraged the development of numerous meandering river systems.
These fed into the great embayment from both the east, where tectonic activity was mild and the
west, where aggressive mountain-building was underway. On land, this major period of marine
transgression brought with it a more humid, temperate climate and a rise in base level. While
thick sequences of organic-rich marine shale accumulated in the axis of this seaway, fluvial
complexes, whose sands would later serve as updip traps for petroleum, fed into its margins.
Continued tectonism divided much of the western part of this seaway into variously-sized
intermontane basins (Denver-Julesberg basin, Laramie basin, Powder River basin, Big Horn
basin, Wind River basin), which individually generated substantial amounts of hydrocarbon
reserves. Due to the rising uplands to the west, many of the river systems were characterized by
higher slopes and thus straighter, less sinuous channels than some modern-day classic rivers
such as the Mississippi.

For example, Figure 1 (Generalized stream channel pattern in the Lloydminster, Alberta area)
shows the configuration of channels in the Lloydminster field area, on the Alberta-Saskatchewan
border. Abundant well data has indicated the pattern shown. The sands themselves are
discontinuous, with individual bodies being less than 30 m thick and 10 km long. The vein-like
pattern displays several generations of stacked channels and should not be mistaken as
indicating braided stream flow. Figure 2 and Figure 3 (Maps delineating small field area in
Lloydminster, Alberta. Part a shows an isopach map of several channels from the
producing area and part b is a structural contour map of the same area. Note the small anticlinal
arch in the westernmost channel sand body) reveals the actual dendritic plan view. A cross
section through one of the sand bodies is given in Figure 4 . (Cross section of fluvial channel
sands in the mannville group, Lloydminster area) and shows the lenticular geometry of the fluvial
sandstones (compare also Figure 5 , Isopach map and cross section of the Cretaceous Denkman
and Little Creek field, Mississippi. Note the scale of the meander loops and the lenticular
nature of the sand body.) Note how the changes in the SP log particularly have been used to
trace the lateral and vertical extent of this channel and its flanking sands.
In this particular case, the main channel sandstones are draped over a basement high that has
created anticlinal trapping. Note the structural closure in Figure 2 and Figure 3 .
As a second example, the Coyote Creek field in the eastern Powder River basin (Wyoming)
provides an excellent example of a point-bar sand reservoir. ( Figure 6 , Isopach map of
Coyote Creek field, Wyoming-eastern Powder River basin and Figure 7 , Log cross section of
point-bar stratigraphic trap in the Cretaceous Fall River Sandstone, also from Coyote Creek field,
Wyoming. Note the bell-shaped gamma ray log curves through the point bar and the subdued
profile through the oxbow fill.)
Isopaching reveals the orientation and large size of the bar, as well as the shale-filled abandoned
(oxbow) channel that forms its hem. Several low areas, called swales, exist on the point bar itself,
and are filled with siltier, low-porosity material. These limit lateral fluid flow. As in the above
example, SP logs provide a relatively reliable indication of the location and extent of the
sandbody. Well A-21 shows a well-developed, bell-shaped curve in the point bar. How does the
curve in this well compare with that shown for Well B-2? The reservoir sands themselves are
rather fine-grained, well-sorted, with porosities between 15% and 24% and permeabilities of up to
300 md. Gross pay is about 40 ft, and total reserves have been calculated at 20 million bbl of oil
(Berg 1968).
1.5 Eolian Environments

General Introduction

Most of the earth's desert regions are in a state of approximate depositional equilibrium, i.e.,
where sedimentation and erosion command equal influence. Fully 25% to 30% of the planet's
surface is arid or semiarid; most of this, however, is composed of bedrock plains, gravel
pavement, and flat sand areas. Only about one-quarter is covered by eolian (wind-blown) sand
deposits, which require a high supply of detritus and sufficient wind velocity to move it. Often, as
shown in Figure 1 (Idealized block diagram of a generic desert environment), this means close
association with environments characterized by coastal orogenic activity, such as alluvial fans,
braided streams, and beaches (e.g., Western North and South America, Mexico) or vast
peneplain regions that act as catch areas for sand transported by very strong, perennial winds
(central Mongolia, Arabia, Northern Africa, Central Australia).

When wind blows across a desert, silt and clay are carried higher up into the atmosphere and
ultimately may be transported great distances to the oceans. Dust that does resettle in the desert
may be trapped and form beds in ephemeral lakes, such as playas. Sand, by contrast, is
transported close to the ground largely by the "leaping" process, which geologists call saltation.
This is triggered partly by hydrodynamic lift and partly by energy transfer from grain collisions.
Wind velocities are seldom sufficiently high to transport gravel.

The bed forms that sand settles into when transported by wind are mainly asymmetric ripples and
dunes whose overall geometry is much like that of their subaqueous counterparts. ( Figure 2 ,
Four main types of sand dunes, as identified from modern deserts. Only transverse and, less
significantly, barchan dunes appear to be represented in the sedimentary record. Seif and stellate
forms seem to represent environments of equilibrium, where sand is moved and reworked, but no
important net sedimentation occurs.) The dynamics of eolian and aqueous movement are
basically similar: they both involve granular solids being moved by and within "fluids." This is most
likely the main reason why the eolian environment is particularly difficult to distinguish in the
subsurface.

As with alluvial fan and braided stream facies, preserved eolian sequences are best known in
rocks of the Late Paleozoic Mid-Mesozoic age, with the Permain period having been especially
hospitable to the development of large coastal sand seas. At the same time, northern Africa, by
virtue of having occupied approximately the same latitude for nearly all of the Phanerozoic, shows
a much more extensive time of eolian deposition.
Depositional Processes and Lithologic Characteristics

Figure 1 (Relationships between dune morphology and the orientation of cross-bedding. Far right
diagrams show azimuth frequency. ) illustrates the four most common modern dune forms.
Discerning specific dune types in the subsurface is only rarely important to the petroleum
geologist. More often, simply deciding whether a facies is actually eolian or not consumes the
major effort; however, a basic understanding of the different types does provide information
essential to the knowledge of what reservoir characteristics can normally be expected.

Barchans and transverse dunes today are transient, unstabilized features that migrate
continuously downwind. Despite this, the crossbed pattern they produce appears typical for the
majority of dune complexes preserved in the sedimentary record. ( Figure 2 , Cross section of a
barchan or transverse dune showing the various bedforms and slipface surfaces. ) In contrast,
stellate (spider-shaped) and seif (knife-edge or longitudinal) dunes, with their more complex
stratification patterns, are permanent features today in the great Saharan, Mongolian, Australian,
and Arabian sand seas, but have not often been identified in lithified formations.
The usual vertical sequence of textures and sedimentary structures seen in most dunes is shown
in Figure 2 . The common subfacies of the interdune area include:

• clay and silt (ephemeral lakes)

• evaporites (salt/gypsum lenses from lake evaporation)

• sand and gravel (as "deflation" lag after the wind has removed smaller detritus)

Crossbed foresets (both tabular and trough types) are often larger (20 ft to 100 ft), steeper (20° to 35°), and
more concave-upward, than in aqueous sands.

Medium-to giant-scale, cross-stratified dune units are commonly truncated — with the low-angle
toeset beds of the overlying unit cutting off the high angle foreset beds of the underlying unit —
and stacked. Such multiple-stacked complexes make up the greater part of eolian formations. (
Figure 3 , Photo of giant scale cross-bedding in the Navajo Sandstone of southern Utah)

Sand grains can show a range of size and rounding, as well as certain surface characteristics,
such as frosting and pitting. The high degree of rounding found in those sands composing the
great ergs ("seas") of north Africa and Arabia is primarily the result of repeated reworking, i.e.,
several generations of erosion and redeposition. In contrast, grains composing the landforms at
Great Sand Dunes National Monument in Colorado, show diversity of specific sedimentologic
traits, including both high-angularity and compositional heterogeneity.
Wind is generally an excellent sifting agent. Though grain size will vary according to position on a
dune, the sorting within each size range is usually good to excellent. Grain orientation can occur
within single crossbed layers, and this can impose a certain amount of directional permeability to
the rock. This will be discussed below in more detail, as it can be an important reservoir
characteristic.

Some dunes, particularly along coast lines, are composed of marine skeletal carbonate sands.
These, however, are rare in the subsurface. Quartz sand dunes also form on barrier islands and
spits, and may thus be incorporated as local subfacies within otherwise transitional or marine
facies.

Like other subaerial sands, eolian complexes have played host to the postburial percolation of
acidic meteoric waters, which have sometimes imparted reddish coloration. Clean, white
sequences, however, are equally common.

Paleontology
Desert dune formations composed almost exclusively of quartz sand are, to a large extent,
predictably barren of fossil remains. Root let horizons, animal tracks, mudcracks, and raindrop
impressions are all evident in present-day wet interdune areas, and are occasionally seen in
outcrop. They are, however, inevitably rare in core samples. Reptile bones, freshwater mollusks,
crustaceans, and arthropods have been reported from the major Late Paleozoic Early Mesozoic
eolianite formations of the western United States, for example:

• Navajo sandstone (Triassic-Jurassic)

• Coconino sandstone (Permian)

• DeChelley sandstone (Permian)

• Lyons formation (Permian)

• Weber-Tensleep sandstone (Pennsylvanian-Permian)

In general, the difficulty of distinguishing eolian complexes points out the importance of outcrop
study for the geologist. Examining rocks in their most exposed state is, of course, an idea
circumstance for most petroleum geologists who must often remain satisfied with the more-or-
less isolated and, at times, random sampling in the form of cores. Let it be said, then, that where
outcrops exist, they should never be passed over in favor of borehole data. Cores are, in a sense,
artificial exposures; the second most direct form of evidence for analysis.

Geometry
There is no distinctive shape to eolian formations. The original dune morphology of a sand sea is
generally reworked and planed out by transgressing waters before another facies is deposited.
Basically, the final shape is a blanket. Thicknesses may be 300 m or more (Navajo sandstone),
but are commonly less. Aerial distribution of both ancient-and modern examples is highly
variable, ranging from tens to hundreds of thousands of square kilometers.

Associated Facies
Generally speaking, desert dune areas are determined by a limited range of physical parameters,
most of which do not involve the depositional variations introduced by water. For this reason,
subfacies in this environment are localized and usually minor. In addition, the often regional
continuity of sand means that facies changes are large-scale. Frequently, adjacent environments
are thickly evaporitic, whether coastal (wadisabkha) or inland (large salt lakes). Associated desert
facies have already been mentioned; they often include alluvial fans, bajadas, playas, braided
streams. ( Figure 1 , Idealized block diagram of a generic desert environment.) The occurrence of
sand dunes on barrier islands and spits has also been discussed. Today, a number of dune
complexes are trimmed by rivers (which cannot be overcome by dune migration) and their alluvial
plains.

In addition to these common associations, eolianites frequently occur immediately above a major
unconformity, or as one of several facies sandwiched within a series of more minor erosion
surfaces marking an episode of fluctuation between sandy marine and nonmarine environments.
( Figure 2 and Figure 3 , Eolian depositional regime showing complex stratigraphic relations
between dune, interdune, playa, and alluvial deposits. ) This desert coast, marginal marine
depositional scheme is apparently responsible for the subaerial/subaqueous confusion and
controversy surrounding a number of well-known stratigraphic intervals in the western United
States. Most of these have been mentioned above and are traditional hydrocarbon reservoirs for
large parts of the Rocky Mountain region. To some degree, the difficulties and controversies
involving these rocks have shown that the definitive separation of eolian and marine facies
becomes at some point an academic exercise (Ahlbrandt and Fryberger 1982).

Diagnostic Evidence

Seismic

In general, subsurface dune deposits are not detectable as such by existing seismic methods.
Sheet-like geometry, association with unconformities, and absence of good internal reflectors are,
as mentioned, also typical of the overall response generated by braided stream sediments (see
previous discussion), which may over or underlie eolianites and thus further "mask" them.
Seismic data, therefore, are perhaps most useful in delineating the depositional limits, rather than
the actual lithology, of a potential dune reservoir.

Wireline Logs
Figure 1 (Log motifs for eolian sands. Note the well-developed blue pattern of upward-
increasing dips along the toeset-foreset transition in individual dune units.) shows a suite of logs
typical of the eolian Rotliegendes (Permian) group, a productive reservoir in the North Sea.
Despite an overall blocky appearance, the gamma-ray curve can be divided into approximately
50-foot increments, all bordered by narrow spikes of higher radioactivity. Thus, the general profile
can be more accurately described as "saw-toothed." Each of the small kicks (which are more
obvious on the density log curve) is asymmetric, with a sharp base and a gentle upward decrease
in API units. They are caused by the finer-grained, mica-containing toeset layers of each new
dune that abruptly truncate the underlying foresets of the underlying dune unit. The wind blows
this finergrained sediment farther out from the dune crest. It thus settles at the base of the dune,
while the coarser quartz grains remain along the slip face. It is this particle size differentiation that
gives the strong concave-upward curve to eolian crossbeds. Note how the kicks on the gamma
ray curve correspond with both the neutron-density and dipmeter logs.
The 50-foot interval is also strongly evident on the latter. Each increment begins at the base with
low angle dips (toeset beds), which then increase in amount upwards until reaching a maximum
of about 25° to 35° (foreset beds). This maximum is the most conspicuous part of the dipmeter
log and indicates both the large size and consistent orientation of the cross-bedding. Dip
azimuths are very constant, directly indicating the downwind direction. This, in turn, reveals the
local elongation of the sand body. As Figure 2 (Relationships between dune morphology and the
orientation of cross-bedding. Far right diagrams show azimuth frequency.) shows, this
elongation will be perpendicular to the downwind direction. Between the stacked dune units, as
shown in Figure 1 , this direction will vary considerably; thus accurate determinations of the
principle paleocurrents can only be made from a relatively large number of readings. Modern day
sand 'seas" are often characterized by two or more major wind directions (Ahlbrandt and
Fryberger 1982). There are only a few, isolated parts of the log that appear to show the low angle
green, or structural, dip motif. Again, the most salient feature of dipmeter logs in this facies is the
pronounced domination by foreset bedding, with its often-steep and highly consistent readings.

Cores

Samples of eolianite sections are commonly comprised almost entirely of clean, well-sorted
quartz sandstone (often called orthoquartzite). Detailed sedimentological analysis — which
examines such aspects as grain mineralogy, size, sorting, rounding, as well as surface
characteristics through both microscopic and mathematical techniques — has not proven its
unqualified worth in strictly distinguishing dunes from some transitional marine facies. Eolian sand
is often reworked from older deposits, and thus such study may reveal mostly "inherited" features.
This, however, is only a cautionary note; such analysis is absolutely necessary to almost all
facies identification — for the geologist, nothing can replace or supersede first-hand examination
of the rocks themselves.

As indicated by the dip log, steep and, at times, monotonously consistent cross-bedding is the
prominent sedimentary structure found in most cores. In addition, smaller-scale ripples, graded
beds, evenly laminated beds, and coarser pebble lag layers along surfaces that divide dune sets
are sometimes seen. Traces of oxidized "impurities" between sand grains — whether ferric iron,
spores, heavy minerals — can be significantly diagnostic. Certainly, as with alluvial fan and
braided stream deposits, coloration can be used (with appropriate caution) as evidence.

The surface features which might be seen in cores, e.g., desiccation cracks, raindrop
impressions, etc., also exist in the braided stream environment and cannot be considered
diagnostic. Alone, core samples will rarely determine whether a deposit is eolian or not. More
often, they must be used in conjunction with detailed log and stratigraphic information.

Reservoir Characteristics
Eolianite sequences are commonly thought to offer some of the best reservoir characteristics.
Though often true, this should not be a reflexive assumption. Among several misconceptions
about this facies is the notion that they are, by definition, very well-sorted, lithologically
homogeneous sandstones with excellent and vertically continuous porosities and permeabilities.

In fact, as discussed by Ahlbrandt and Fryberger (1982), dune sequences are known to be
heterogeneous, complex reservoirs with substantial vertical and lateral variation in fluid
conductivity. One reason for this is the differential texture that exists along individual crossbeds,
which essentially influences differential fluid flow. Specifically, grain orientation normally takes
place along each cross-bed, with micas and more lenticular sand grains aligned in the toeset
plane. In addition, toeset beds often show alterations of such finer-grained material with much
coarser sand that has avalanched down the slipface of the dune. These aspects introduce
significant permeability heterogeneity. Moreover, cementation is also apparently influenced by
these anisotropies and, therefore, differential porosity/permeability preservation has also resulted
in some deposits. As a result, fluids may be best able to migrate along, rather than across,
crossbed layers, i.e., horizontal permeabilities commonly exceed vertical permeabilities. Figure 1
(Porosity/permeability plot showing variety in general reservoir characteristics for the productive
eolian Leman sands of the North Sea region) is a simple indication of the wide and continuous
range in reservoir quality shown by eolianites.

In terms of known traps, structural, stratigraphic, and combination situations are well-studied.
The Nugget sandstone, for example, is now widely accepted as being mainly eolian, and forms
one of the major reservoirs along the Western Overthrust Belt of the Rocky Mountain region.
Figure 2 (Northwest-southeast cross section through Painter Reservoir field showing complex
structure typical along the western Overthrust. Note that the trap is still a relatively coherent
anticline.) and ( Figure 3 , Isopach map and discovery well log (gamma ray/neutron/density) of
Painter Reservoir field.for isopaching the net effective reservoir, 10% minimum porosity was
used) shows an instance of the complex structure that characterizes Nugget fields along the
producing fairway in southwesternmost Wyoming.
More generally, eolianites share with alluvial fan and braided stream facies the potential
drawback of being deposited at considerable distances from marine source rocks. Certainly, the
"telescoping" together of entire stratigraphic sections which takes place during regional thrust
faulting ( Figure 2 ) is one way nature has used to juxtapose widely disparate facies. At the same
time, dune deposition does, in fact, occur quite frequently in proximity to coastlines. Furthermore,
the nature of this deposition reduces the opportunity for subfacies development and, therefore,
large-scale stratigraphic traps are possible, for example, at the updip margins of marine basins.
Regional pinch-out into, or capping by, evaporitic facies is relatively common in the sedimentary
record, as shown by the Rotliegendes reservoirs of the North Sea.

Examples

The Rotliegendes (Lower Permian) group of the southern North Sea basin contains a series of
giant natural gas fields, including the onshore Groningen field (Netherlands), and the offshore
West Sole, Viking, Leman, and Indefatigable fields. ( Figure 1 , Location of the Viking gas field
area in the southern North Sea region.) Proved and probable reserves combined for this play are
on the order of 85 tcf. Though the Rotliegendes itself contains a diversity of continental
deposits (including fanglomeratic sands, braided stream/wadi conglomerates, playa/sabkha silts
and shales), the most favorable reservoir is decidedly the eolian Leman sandstone.
In this region of northwestern Europe, the Late Paleozoic saw the development of an orogenic
belt that stretched from what is today southern Ireland, across southern England and the low
countries into Germany. An arid climatic region existed at that time and a suite of desert facies,
including alluvial fan, braided stream, eolian dune, coastal sabkha, and evaporitic marine
environments, developed from the rising highlands northwards into a subsiding southern North
Sea basin. ( Figure 2 , Distribution of environments and facies in the Rotliegendes group. See
Figure 3 , A schematic cross section through the Trucial coast of Oman, for comparison. ) The
basin was bounded to the north by a low ridge, now called the mid-North Sea high. Prior to basin
subsidence, the region had undergone a period of erosion, which truncated gently folded
Pennsylvanian Coal Measures. These were buried by the Rotliengendes redbeds which, in turn,
were overlain and "sealed" by the carbonates and evaporites of the Upper Permian Zechstein
when the ocean breached the mid-North Sea high. Subsequent tectonism broke up the basin
floor into horsts and grabens, and the increased pressure/temperature conditions associated with
this movement helped generate gas in the buried coals. The Rotliegendes were thus in an ideal
position to act as reservoirs for this gas.

As shown in Figure 4 , (Map view of the facies progression shown in Fig.2, as interpreted for the
Viking gas field are. ) three main facies characterize the Rotliegendes of the producing area.
These were identified largely on the basis of core samples (Gage 1980), which served as a basic
guide for log analysis. Along the southern flank of the basin adjacent to the mountain front are red
conglomerates and coarse, pebbly, poorly sorted, crossbedded sands, interpreted as the remains
of a braided wadi outwash plain, with the most proximal, fanglomeratic facies believed to indicate
fan deposition.

Due to poor sorting, this facies has low to negligible reservoir quality. It was deposited by a series
of ephemeral rivers flowing from west to east, out of the highland area and through a region of
eolian dunes. These fluvial deposits inferfinger to the north and south with medium to fine-
grained, well-sorted sands which, as shown by core and dipmeter work, contain giant-scale,
concave-upward crossbeds 50 ft high or more. Logs through this facies are very similar to that
shown in Figure 5 . (Log motifs for eolian sands. Note the well-developed blue pattern of
upward-increasing dips along the toeset-foreset transition in individual dune units.) This
dominantly eolian facies grades northwards into red shales and evaporites, which suggest
deposition in playa lake and sabkha environments. Grain sorting appears to be the main factor
affecting reservoir quality in all facies.
Rotliegendes reservoir characteristics are, as indicated, directly tied to facies.
( Figure 6 , Overall relationships between facies and permeability, Rotliegendes Group, Viking
gas field area and Figure 7 , Porosity/permeability plot for the various facies in the Viking gas field
area. ) Their variability necessitated detailed mapping of basin paleogeography in stages.
These maps were able to show how a series of specific alluvial fans prograded northwards
towards an enormous playa lake. The sediments of these fans were winnowed out and reworked
by desert winds which then redeposited them in well-sorted form as a surrounding field of dunes.
Such mapping was not done simply to reconstruct picturesque deserts of the past. In the case of
the Rotliegendes, it was able to not only explain, but to predict reservoir heterogeneities,
thicknesses, and lateral extent. This example further reveals that not all braided stream deposits
are alike; they do not universally possess high reservoir quality. Facies analysis, then, is most
often just the beginning to cogent, useful knowledge about a particular basin or formation.

Exercise

Refer to logs, A, B, and C ( Figure 1 , Gamma ray and resistivity logs for wells A,B,C ) and their
locations on the accompanying map Figure 2 .
General lithologic descriptions from cuttings samples indicate the following top-to-bottom rock
types:

Well A: Shale with thin sand interbeds. Coarse-medium grained, cross bedded sand, some gravel,
thin shale interbeds. Coarse conglomerate, interbeds of cross bedded sand/gravel, thin shale
interbeds. Coarse-medium grained, cross bedded sand, shale interbeds.

Well B: Same as A, conglomeratic unit thinner.

Well C: Shale with thin sand interbeds. Coarse-medium grained, cross bedded
sand/gravel, thin shale interbeds; sands show rippling, rip-up clasts of shale.

(a) Match rock descriptions with portions of the log curves and identify the probable facies. High-angle
reverse faults trending NW-SE and dipping to the NE are known to occur in the area.

(b) Sketch in the facies boundaries between logs A, B, and C, showing where you interpret any
faults to exist.

(c) Sketch a generalized porosity vs. depth curve for each well.

(d) Referring to the accompanying location map, logs from Wells D and E are very similar to
those in Well A, but facies contacts appear to occur 100 ft deeper. Based on this, what would you
expect the lithology and log curves to be in Well X?

(e) Sketch on the map a generalized paleogeographic interpretation.

Include paleocurrent arrows.

(f) Good shows of oil occur in all three wells at about 850-900 ft depth. Based on your above
interpretations, roughly where would your best prospect location be?

Solution
Figure 3 and Figure 4 (Interpretation of well log cross section; generalized porosity profiles for
each well.)

The three basic facies indicated by the well logs and lithologic descriptions are shale (possibly
meandering stream flood plain), braided stream, and alluvial fan. Sawtooth portions of the gamma
ray curves indicate the thickness of the fan facies contained within the more blocky profile
generated by the braided alluvium. The fault shown on the map may be assumed to be
genetically related to the tectonic front which exists to the southwest. Structurally, it would appear
that the uplift associated with this front occurred as a rather sudden pulse, generating a single
alluvial fan which, for a short time, lapped over braided stream deposits. Sediment supply is from
this uplift, i.e., from southwest to northeast. The mapped fault, which parallels this front, has
created the local high in the area of Wells A and B.

Figure 5 (Interpreted paleography. )

The lithology in well X would then be dominated by fanglomerate, and the corresponding log
curve would show a high degree of serration throughout. A good, though rough prospect location
would be between Wells A and B, on the upthrown side of the fault. Drilling would penetrate into
the downthrown side, where both the fault and the relatively impermeable fan deposits would
have conspired to form a trap. Porosity would be fairly low within the fan facies due to the content
of clay that normally occurs within fanglomerate.

RECOMMENDED READING
The following books and papers are recommended for general background reading:

Ahlbrandt, T.S., and S.G. Fryberger, 1981, Sedimentary features and significance of interdune
deposits, in: Recent and Ancient Non-Marine Depositional Environments: Models for Exploration,
F.G. Ethridge and R.M. Flores, eds., Society of Economic Paleontologists and Mineralogists,
Special Publication No. 31., pp. 293-314.

Balducchi, A., and G. Pommier, 1970, Cambrian Oil Field of Hassi Messaoud, in: Geology of
Giant Petroleum Fields, M.T. Halbouty, ed., AAPG Memoir 14., pp. 477-488.

Bennacef, A., S. Beub, B. Biju-Duval, O. De Charpal, O. Gariel, and P. Rognon, 1971, Example
of Cratonic Sedimentation: Lower Paleozoic of Algerian Sahara, AAPG Bull., vol. 55, pp. 225-
2245.

Chatfield, J., 1982, Case History of Red Wash Field, Uinta County, Utah, in: Stratigraphic Oil and
Gas Fields, R.E. King, ed., AAPG Memoir 16., pp. 342-353.

Chin Chen, 1980, Non-Marine Setting of Petroleum in the Songliao Basin of Northeastern China,
Journal of Petroleum Geology, vol. 2, p. 233-264.

Cole, R.D., and M.D. Picard, 1981, Sulfur-Isotope Variations in Marginal Lacustrine Rocks of the
Green River Formation, Colorado, and Utah, in: Recent and Ancient Non-Marine Depositional
Environments: Models for Exploration, F.G. Ethridge and R.M. Flores, eds., Society of
Paleontologists and Mineralogists, Special Publication No. 31, pp. 261-275.

Collinson, J.D., and S. Lewin, 1983, Modern and Ancient Flu vial Systems, International
Association of Sedimentologists, Special Publication No. 6, p. 575. This volume contains 44
papers on recent and ancient fluvial deposits, including one or two that deal with petroleum
reservoirs.

Ethridge, F.G., and R.M. Flores, 1981, Recent and Ancient Non-Marine Depositional
Environments: Models for Exploration, Society of Economic Paleontologists and Mineralogists
Special Publication No. 31, p. 349. This volume contains papers covering fluvial, lacustrine and
eolian deposits, including several dealing with petroleum.

Evans, P.R., 1981, The Petroleum Potential of Australia, Journal of Petroleum Geology, vol. 4,
pp. 123-146.

Glennie, K.W., 1970, Desert Sedimentary Environments, Elsevier Scientific Publishing Company
Inc., Amsterdam, p. 222. The definitive textbook on desert environments.

Glennie, K.W., and B.D. Evamy, 1968, Dikaka: Plants and Plant Root Structures Associated with
Eolian Sand, Paleogeography, Paleoclimatology, Paleoecology, vol. 4, pp. 77-87.

Grunau, H.R., and U. Gruner, 1978, Source Rock and Origin of Natural Gas in the Far East,
Journal of Petroleum Geology, vol. 1, pp. 3-56.

Guangming, Z., and Z. Ouanheng, 1982, Buried-Hill Oil and Gas Pools in the North China Basin,
in: The Deliberate Search for the Subtle Trap, M.T. Halbouty, ed., AAPG Memoir 32, pp. 317-336.

Harms, J.C., P. Tackengerg, E. Pickles, and R.E. Pollock, 1981, The Brae Ollfield Area, in:
Petroleum Geology of the Continental Shelf of Northwest Europe, L.V. Illing, & G.D. Hobson,
eds., Heydon Press, London, pp. 352-357.
Jones, H.P., and R.G. Speers, 1976, Permo-Triassic Reservoirs on Prudhoe Bay Field, North
Slope, Alaska, in: North American Oil and Gas Fields, J.E. Baunstein, ed., AAPG Memoir 24, pp.
23-50.

Jopling, A.V. and C.B. McDonald, eds., 1975, Glaciafluvial and Glaciomarine Sedimentation,
Society of Economic Paleontologists and Mineralogists, Special Publication No. 23, p. 320.

Keighin, C.W., and T.D. Fouch, 1981, Depositional Environments and Diagenesis of Some Non-
Marine Upper Cretaceous Reservoir Rocks, Uinta Basin, Utah, in: Recent and Ancient Non-
Marine Depositional Environments: Models for Exploration, F.G. Ethridge, and R.M. Flores, eds.,
Society of Economic Paleontologists and Mineralogists, Special Publication No. 31, pp. 109-125.

Leopold, L.B., M.G. Woman, and J.P. Meller, 1964, Flu vial Processes in Geomorphology: W.H. -
Freeman and Co., San Francisco, p. 522.

Levandowski, D.W., M.E. Kaley, S.R. Silverman, and R.G. Smalley, 1973, Cementation in Lyons
Sandstone, Role in Oil Accumulation, Denver Basin, Colorado, AAPG Bull., vol. 57, pp. 2217-
2244.

Marie, J.P., 1975, Rotliegendes Stratigraphy and Diagenesis, in: Petroleum and the Continental
Shelf of Northwest Europe, A.W. Woodland, ed., Applied Science Publishers Ltd., London, pp.
205-212.

McGregor, A.A. and C.A. Biggs, 1970, Bell Creek Oil Field, Montana, in: Stratigraphic Oil and
Gas Fields, R.E. King, ed., AAPG Memoir 16, pp. 367-675.

McKee, E.D., ed., 1979, A Study of Global Sand Seas, U.S. Geological Survey, Special Paper
1052, 427 p. A well-illustrated review of recent eolian sands and their probable ancient analogs.

Moody, S.M., 1966, High and Low Sinuosity Stream Deposits with Examples from The Devonian
of Spitsbergen, Journal of Sediment Petroleum, vol. 36, pp. 1102-1117.

Morgride, D.L., and W.B. Smith, 1972, Geology and Discovery of Prudhoe Bay Field, Eastern
Arctic Slope, Alaska, in: Stratigraphic Oil and Gas Fields, R.E. King, ed., AAPG Memoir 16, pp.
489-501.

Randy, R.R., 1982, Seismic Stratigraphic Interpretation of the Fort Union Formation, Western
Wind River Basin: Example of Subtle Trap Exploration in a Non-Marine Sequence, in: The
Deliberate Search for the Subtle Trap, M.T. Halbouty, ed., AAPG Memoir 32, pp. 169-180.

Reineck, H.E., and l.B. Singh, 1973, Depositional Sedimentary Environments, Springer-Verlag,
New York.

Rigby, J.K., and W.K. Hamblin, eds., 1972, Recognition of Ancient Sedimentary Environments,
Society of Economic Paleontologists and Mineralogists, Special Publication No. 16, 340 p.

Rust, B.R., 1979, Fades Model 2, Coarse Alluvial Deposits, in: Facies Models, R.G. Walker ed.,
Geoscience Reprint Series 1, pp. 9-21

Scholle, P.A., and D. Spearing, eds., 1982, Sandstone Depositional Environments, AAPG, Tulsa,
410 p. A well-illustrated series of papers on fluvial, lacustrine, and eolian deposits, including
examples that are petroleum reservoirs.
Schram, M.W., E.V. Dedham, and J.P. Lindsey, 1977, Practical Stratigraphic Modelling and
Interpretation,in: Seismic Stratigraphy — Applications to Hydrocarbon Exploration, Payton, ed.,
AAPG Memoir 26, pp. 477-502.

Schumm, S.A., 1972, Fluvial paleochannels, in: Recognition of Ancient Sedimentary


Environments, Society of Economic Paleontologists and Mineralogists, Special Publication No.
16, pp. 98-107.

_________ 1981, Evolution and Response of the Fluvial System, Sedimentological Implications,
in: Recent and Ancient Non-Marine Depositional Environments: Models for Exploration, F.G.
Ethridge, and R.M. Flores, eds., Society of Economic Paleontologists and Mineralogists, Special
Publication No. 31, pp. 19-30.

Selley, R.C., 1969, Torridonian Alluvium and Ouicksands, Scottish Journal of Geology, vol. 5, pp.
328-346.

_________ ,1972, Diagnosis of Marine and Non-Marine Environments from the Cambro-
Ordovician Sandstones of Jordan, Journal of the Geological Society, vol. 128, pp. 135-150.

Smith, D.G., 1983, Anastomosed Fluvial Deposits: Modern Examples from Western Canada, in:
Modern and Ancient Fluvial Systems, J.D. Collinson and J. Lewin, eds., International Association
of Sedimentologists, Special Publication No. 6, pp. 155-168.

Tanner, W.F., 1967, Ripple Mark Indices and Their Use, Sedimentology, vol. 9, pp. 89-104.

Visher, G.S., 1971, Depositional Processes and the Navajo Sandstone, Geological Society of
America Bull., vol. 82, pp. 1421-1424.
2. MARGINAL MARINE SANDSTONE RESERVOIRS

Major Depositional Influences in Marginal Marine Systems


The basic influences that determine clastic depositional patterns in marginal marine settings can
be divided into those of nonmarine and marine origin. Both of these can, in turn, be influenced by
climate and tectonics.

Nonmarine parameters include the following:

• overall river discharge

• sediment supply

• initial texture and composition

Major marine influences encompass


• waves

• tides

• near-bottom currents

Climate controls a number of the parameters mentioned above. In particular, rate of erosion, river
discharge, and thus sediment supply are all directly tied to climatic effects. In addition, seasonal
storms add sudden and significant amounts of energy to an environment. Such energy can
redistribute large amounts of sediment, both along the coastline as well as on the shelf bottom at
depths below the normal fair weather wave base.

Finally, tectonic factors can have an important effect on the composition, supply, and distribution
of sediment. Depositional slope will directly influence the location and geometry of marginal
marine sand bodies. Steeper slopes tend to restrict coastal deposition and enhance the transport
of material into deeper water. Conversely, lower depositional slopes allow marginal marine
environments to become laterally extensive and continuous.

Other tectonic factors include coastal uplift or subsidence. High coastal mountain ranges or areas
of active uplift usually prevent the development of continuous, sand-dominated shorelines. Low
coastal plain relief and gradual subsidence, on the other hand, provide an excellent setting for
marginal marine environments to become fully developed and to migrate. Migration can be either
landward (transgressive), if subsidence is rapid enough, or basinward (regressive), if sediment
supply is high enough.

High sediment influx is often associated with passive continental margins where drainage divides
are far from coastlines, and river systems discharging sediment to the coast have large drainage
areas. On the other hand, drainage areas are usually limited along active continental margins,
and sediment influx tends to be correspondingly low.
Most marginal marine depositional systems of importance to petroleum exploration (deltas and
barrier islands) are currently being deposited along passive continental margins (e.g., U.S. Gulf
Coast and both margins of the Atlantic Ocean). Maximum preservation is associated with
relatively rapid regressive sedimentation, which acts to bury each subenvironment. Marine
transgression, by contrast, usually causes erosion of the seaward, sandy portion of a coastal
deposit.

Major regressive episodes (with secondary transgressive cycles) have characteristically been
associated with the postrift or drift stages of continental separation. Sedimentation during this
phase is supported by mature river systems with large drainage basins that commonly extend into
igneous and metamorphic shield areas. Such areas provide quartz-rich sand for transport to
marginal marine settings.

On a smaller scale, such regression has also occurred many times along shallow inland seas,
such as those that existed in North America and Europe during the Mesozoic and Middle to Late
Paleozoic.

Figure 1 (Seismic section from offshore Africa showing several major depositional sequences.
Units K1.2 and K1.3 are of dominantly marginal marine origin.) is a seismic profile whose
interpretation is based on projected well control and surface geology. The section, from offshore
northwest Africa, shows the variety of tectonostratigraphic relations common along many passive
margins. In particular, it shows several occurrences of marginal marine facies.

Beginning at the base, we might first note the upward changes in depositional slope that occur
between the Triassic-Jurassic (TR-J1) and the Tertiary section (TP&E, TM2, and so on). The
overall depositional scheme is strongly progradational. Specifically, units J1 through J3.2 are
interpreted as carbonate shelf sediments that built out over a sequence of syn-rift Triassic red
beds, evaporites, and volcanics (Vail et al. 1978). Therefore, they presumably represent the first
regressive episode following marine incursion of the young rift valley.
By Lower Cretaceous time, clastic sediment supply had become pronounced enough to initiate
the second major episode of marginal marine deposition; most of the resulting deposits (units
K1.2, K1.3) are composed of deltaic sediments on the right, prograding over deep marine slope
shales to the left. These accumulated rapidly and loaded the shelf, producing the series of
growth-type faults shown. Overlying this sequence is unit K2, which is interpreted as a marine
shelf deposit. The Tertiary section is a group of marine shale units that form a set of prograding
and onlapping sequences.

Overall, marginal marine facies occupy approximately 50% to 60% of the total sedimentary
section seen in Figure 1 . These facies include sand-dominated intervals that represent
deposition in environments such as deltaic distributary channels and channel mouths, barrier
islands, and coastal beaches. Such intervals are often characterized by excellent reservoir quality
and, at the same time, are bounded both laterally and vertically by fine-grained, sealing
lithologies.

In areas such as the U.S. Gulf Coast, where rates of clastic sediment supply are relatively rapid
and continuous, carbonate deposition is much more limited and the overall percentage of
marginal marine facies, particularly deltaic facies, can be much higher than that shown in Figure 1
. Generally, both structural traps (e.g., those produced by growth faulting) and stratigraphic traps
(e.g., those resulting from lateral pinch-out) should be common within such large, regressive
depositional wedges.

Additionally, many petroleum reservoirs have been discovered in barrier island sand bodies
associated with ancient epeiric seas. Among the best examples are those found in foreland-type
basins of west-central North America, where extensive Late Mesozoic transgressions inundated
large portions of the continent. Sediment was supplied by actively uplifting mountain ranges to the
west as well as exposed crystalline shield areas on the east. A variety of marginal marine
environments developed along the borders of the shallow Jurassic and Cretaceous seaways;
extensive study of these, combined with studies of recent coastlines, has formed the basis for
much of our current understanding concerning barrier island reservoir environments.

To summarize briefly, four principal influences — terrestrial, marine, climatic, and tectonic —
combine to determine the character and distribution of sediment in marginal marine
environments. Knowledge of each of these, as well as the interaction among them, is necessary
for a full understanding of facies trends and diagenetic patterns. Such knowledge, in turn, forms
the basis for predicting reservoir geometry and quality trends.

General Classification of Sand-Depositing Marginal Marine Environments


The basic transitional marine settings in which sand is deposited on the earth today are the
following:

• deltas

• barrier islands

• coastal plain beaches

• estuaries

• tidal flats
These geomorphic features most often occur in combination. For example, barrier islands are
commonly associated with fluvial coastal plain and lagoonal-tidal flat-marsh environments.

The first two environments have proved to be significantly more important to petroleum
exploration. Coastal beach sands closely resemble those deposited on the oceanward side of
barrier islands. Tidal flats are described as an associated facies of barrier islands. Estuaries,
meanwhile, are the drowned mouths of large rivers and exist where the sediment load of the river
is too small to "overcome" subsidence or tidal action. Based on descriptions of their modern
counterparts, ancient estuarine sand bodies should have good oil and gas potential. However,
descriptions of petroleum production from estuarine deposits are, at best, extremely rare.

Principal Sedimentary Processes


Those processes specifically responsible for deposition in transitional marine environments can
be divided into those related to rivers and those common to the ocean.

Rivers entering the sea generally produce a lobate wedge of fresh, relatively low-density water.
This rides and spreads over the denser, saline water of the open ocean or the brackish water of
bays and lagoons. The mixture of fresh and marine water sets up a series of local bottom
currents that act to replenish the diluted salt water. The zone along which the flow of fresh water
leaves the bottom is frequently where the river drops its heaviest bed load as a distributary mouth
bar.

Coarser sediment initially settles out relatively rapidly and therefore close to the river mouth, while
finer material remains in suspension longer and is carried farther seaward. Much of the coarse
material that is too heavy to be transported out to sea is reworked by marine processes.

Marine distribution of sediment is accomplished by the action of wind-driven waves, currents, and
tides. Wave action is directly proportional to wind speed, and it is therefore during stormy periods
that wave action is most effective. As waves break against the shore they generate a system of
currents: longshore currents flowing parallel to shore and rip currents flowing away from the
shore. ( Figure 1 , Diagram showing the near-shore current system along a typical sandy coast.)
When waves approach the shoreline obliquely, longshore currents are driven in one direction.
Tidal range (i.e., tidal energy) may vary considerably from one coastline to another depending on
a number of factors, which include shoreline geometry and shelf width. In fact, a widely used
classification of coastlines is based on tidal range, as follows: (1) microtidal, where tidal range is
less than 2 m; (2) mesotidal, where tidal range is 2 to 4 m; and (3) macro-tidal, where tidal range
exceeds 4 m (Davies 1964). Hayes (1975), in a study of coastal charts of the world, discovered
that the morphology of a coastline could be related to its tidal classification. Figure 2 (Variation of
morphology of shorelines with respect to tidal range) shows the relative abundance of coastline
depositional features with respect to these various tidal ranges.

Tidal currents affect the distribution of sediment in specific ways. Coastal areas subject to the
direct influence of tidal action are called peritidal (peri- = "encompassing"). Such areas are
divisible into three main zones on the basis of exposure to the atmosphere:

• Subtidal Zone-- This includes the zone below the level of low tide. It is perennially
submerged and is usually subject to gentle tide-related bottom currents. However, these
bottom currents are often sufficiently strong to deposit significant amounts of sand.

• Intertidal Zone-- This is also sometimes referred to as the "littoral zone" and exists
between the levels of high and low tide. It is, therefore, subject to the greatest
concentration and fluctuation of tidal energy, i.e., to diurnal flooding and exposure.

• Supratidal Zone-- This encompasses the area between the levels of normal high tide
and that of the highest storm tide.

The relative height or position of these zones will change over brief periods of geologic time. The
deposition of high-quality reservoir sand mainly occurs in the sub- and intertidal zones.
The interaction between tides, waves, and currents can produce temporary or highly changeable
patterns of sedimentation. Their individual effects, therefore, may be sporadically preserved or
difficult to identify and interpret in a particular sand body. In addition, storms can add large
amounts of energy, causing considerable sediment redistribution in a very short (geologically
"instantaneous") amount of time.

Overall, the diversity of sediment transport energy in the marginal marine setting helps guarantee
a corresponding variation in deposition and, therefore, in grain size, composition, and texture.
The complex interaction of river and ocean means that the lateral and vertical distribution of
facies and subfacies is complex; very different sedimentary regimes, such as lagoons and eolian
dunes, may exist in close proximity. Facies identification and analysis, therefore, can be
particularlychallenging.
3. MARINE SANDSTONE RESERVOIRS
4. DEEP MARINE SANDSTONE RESERVOIRS

Potrebbero piacerti anche