Sunteți pe pagina 1din 532

FACIES MODELS REVISITED

Edited by:
HENRY W. POSAMENTIER
Anadarko Petroleum Corporation, 1201 Lake Robbins Drive, The Woodlands, Texas 77380, U.S.A.
AND
ROGER G. WALKER
Roger Walker Consulting Inc., 83 Scimitar View NW, Calgary, Alberta T3L 2B4, Canada

Copyright 2006 by
SEPM (Society for Sedimentary Geology)

Laura J. Crossey and Donald S. McNeill, Editors of Special Publications


SEPM Special Publication 84

Tulsa, Oklahoma, U.S.A. September, 2006


SEPM (Society for Sedimentary Geology) is an international not-for-profit Society based in Tulsa,
Oklahoma, U.S.A.. Through its network of international members, the Society is dedicated to the
dissemination of scientific information on sedimentology, stratigraphy, paleontology, environmental
sciences, marine geology, hydrogeology, and many additional related specialties.

The Society supports members in their professional objectives by publication of two major scientific journals, the
Journal of Sedimentary Research (JSR) and PALAIOS, in addition to producing technical conferences, short courses,
and Special Publications. Through SEPM's Continuing Education, Publications, Meetings, and other programs,
members can both gain and exchange information pertinent to their geologic specialties.

For more information about SEPM, please visit www.sepm.org.

ISBN 1-56576-121-9
© 2006 by
SEPM (Society for Sedimentary Geology)
6128 E. 38th Street, Suite 308
Tulsa, Oklahoma 74135-5814, U.S.A.
Printed in the United States of America
FACIES MODELS REVISITED

Henry W. Posamentier and Roger G. Walker, Editors

CONTENTS

Facies models revisited: Introduction


ROGER G. WALKER .................................................................................................................................................................................... 1

Eolian facies models


NIGEL P. MOUNTNEY .............................................................................................................................................................................. 19

Fluvial facies models: Recent developments


JOHN S. BRIDGE ......................................................................................................................................................................................... 85

Estuarine and incised-valley facies models


RON BOYD, ROBERT W. DALRYMPLE, AND BRIAN A. ZAITLIN. ................................................................................................ 171

Deltas
JANOK P. BHATTACHARYA ................................................................................................................................................................ 237

A reexamination of facies models for clastic shorelines


H. EDWARD CLIFTON ............................................................................................................................................................................ 293

Facies models revisited: Clastic shelves


JOHN R. SUTER ........................................................................................................................................................................................ 339

Deep-water turbidites and submarine fans


HENRY W. POSAMENTIER AND ROGER G. WALKER ..................................................................................................................... 399

Index ............................................................................................................................................................................................................ 521


FACIES MODELS REVISITED 1

FACIES MODELS REVISITED

ROGER G. WALKER
Roger Walker Consulting Inc., 83 Scimitar View NW, Calgary, Alberta T3L 2B4, Canada
e-mail: walkerrg@telus.net

ABSTRACT: The papers contained on this CD mostly originate from a session at the 2002 Annual Meeting of the Canadian Society of
Petroleum Geologists, repeated at the 2004 Dallas AAPG Meeting. The theme of both sessions was “Facies Models Revisited”, to see what
sort of progress had been made since the third (1992) edition of Facies Models, published by the Geological Society of Canada.
During the ten years between 1992 and 2002, there has been considerable progress in the understanding of modern and ancient
depositional environments. This additional complexity makes modeling much more difficult, and raises the problem of whether
modeling still serves a purpose. The original reasons for creating facies models still exist—a model is a point of comparison, it is a guide
for further observations, it serves as a basis for hydrodynamic interpretation, and most importantly, it acts as a predictor in new
situations.
Using submarine fans as an example, it is clear that progress during the last ten years (particularly in 3-D seismic) has highlighted
the inadequacy of all pre-existing models—indeed, no comprehensive models have been proposed since the mid eighties. Yet with
continued and increasing exploration in submarine fan systems, predictive models are even more necessary. The traditional approach,
of distilling the features that modern and ancient systems have in common, is extremely difficult (and probably naive) in such diverse
and complex systems. Instead, it is necessary to identify all of the constituent building blocks of submarine fans (channels, point bars,
levees, splays, frontal lobes and so on), and try to identify the salient features of each. New models for particular situations can be
constructed by examining the relationships of the constituent building blocks. For example, sinuous channels, levees and splays may
be closely related in space, whereas frontal lobes are unlikely to be related to sinuous leveed channels (except for the channel that
ultimately feeds the lobe). A three-dimensional reconstruction can therefore be made by examining the building blocks that are closely
and commonly related, and also using information from the building blocks that are seldom or never found in juxtaposition.
These principles, discussed above for submarine fans, can be applied to all depositional environments, at all scales. The ideas are
elaborated in this introductory paper, and can be seen in the other contributions to this CD.

INTRODUCTION controlling physical processes of the elements represented in


each environment. The object is to identify the salient features of
“Facies Models”, a publication of the Geological Association recent sediments and ancient rocks, such that these features can
of Canada, first appeared in 1979 (Walker, 1979), with second and be identified, combined, and distilled into models that character-
third editions in 1984 and 1992 respectively (Walker, 1984; Walker ize that particular environment. Once a model is available, how-
and James, 1992). In 2002, the Canadian Society of Petroleum ever simple and basic, it can be used to further our understanding
Geologists organized a session at their annual meeting entitled of natural systems. Perhaps the primary use is in the prediction
“Facies Models Revisited”. The idea was to review progress in and interpretation of sandbodies in oil and gas reservoirs, but
facies modeling during the ten years since publication of the third increasing applications can be found in the movement of ground-
edition of “Facies Models” (Walker and James, 1992). The all-day water through clastic sediments, and in environmental studies.
session was well received, and SEPM requested that a similar Facies models also play an important role in the understanding,
session be organized at the annual AAPG meeting in Dallas (2004). prediction and amelioration of coastal erosion, and Hurricane
This CD includes most of the papers presented in Dallas. We Katrina in 2005 emphasizes the importance of incorporating
have taken advantage of the CD format by including abundant isolated extreme events into the formulation of facies models—
full color illustrations of the examples discussed. The papers on this is a topic that has been underemphasized and is in need of
this CD cover clastic sediments only, and they are more compre- further development.
hensive than the reviews in “Facies Models”. This partly reflects
the advances made during the last ten years, with increasing Facies
recognition of the complexity and variability of depositional
environments. The concept of facies is a very old one, and was introduced
This paper is organized in terms of increasing scale and into geology by Nicholas Steno in 1669. It implied the entire
complexity. The concept of facies will be introduced first, fol- aspect of a part of the earth’s surface during a certain interval of
lowed by facies associations and facies successions. Then the geological time (see Teichert, 1958). The modern usage was
stage is set for a discussion of facies models, with a final discus- introduced by Gressly (1838), implying the sum total of the
sion of future approaches to modeling. lithological and paleontological aspects of a stratigraphic unit.
Translations of Gressly’s extended definition are given by
FACIES AND FACIES MODELING Teichert (1958) and Middleton (1973). The linkage of modern
and ancient environments probably dates back to Johannes
Facies modeling, as understood today, involves a synthesis of Walther in 1893. He suggested that “the most satisfying genetic
information from ancient and recent depositional environments, explanations of ancient phenomena were by analogy with mod-
in an effort to understand the nature, scale, heterogeneity, and ern geological processes” (quoted by Middleton, 1973, p. 981).

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 1–17.
2 ROGER G. WALKER

Facies—A Working Definition this is rather different from the modern usage, where a facies
model is on the scale of individual depositional environments
The most useful modern working definition of the term and may contain many different facies. However, Potter (per-
“facies” was given by Middleton (1978), who noted that: sonal communication, 2005) noted that his facies models were
created “to improve prediction and understanding of how the
“the more common (modern) usage is exemplified by de different lithologies that form a recurring facies are put together”.
Raaf et al. (1965) who subdivided a group of three forma- It is apparent that his “facies” are defined on a larger scale than
tions into a cyclical repetition of a number of facies distin- those discussed in this present paper. Finally, Potter noted that
guished by “lithological, structural and organic aspects “improved prediction largely depends on relating the facies to
detectable in the field”. The facies may be given informal basin geometry and understanding the internal transport system
designations (“Facies A” etc.) or brief descriptive designa- of the facies; together both help us understand the ‘fabric’ of the
tions (“laminated siltstone facies”) and it is understood facies”.
that they are units that will ultimately be given an environ- Despite the very forward-looking ideas expressed at Potter’s
mental interpretation; but the facies definition is itself conference, the term “facies model” did not catch on, perhaps
quite objective and based on the total field aspect of the because there was insufficient information regarding deposi-
rocks themselves… . The key to the interpretation of facies tional environments and architectures to make much progress.
is to combine observations made on their spatial relations The term was reintroduced in 1975 in my paper “Generalized
and internal characteristics (lithology and sedimentary facies models for resedimented conglomerates of turbidite asso-
structures) with comparative information from other well- ciation” (Walker, 1975). At the time, I was unaware of Potter’s
studied stratigraphic units, and particularly from studies earlier usage of the term. Shortly afterward I used the term again
of modern sedimentary environments”. in the first edition of Facies Models (Walker, 1979), not realizing
that the term would become so widely used in sedimentary
The term facies can be used in both a descriptive and an geology. In the second edition of Facies Models (Walker, 1984), it
interpretive sense. The definition above defines only the descrip- is clear that “facies” are the smaller-scale building blocks of the
tive facies. However, it may be useful, as a quick means of sedimentary record. Facies can be organized into “facies se-
communication, to tell a friend that you worked on a “fluvial quences” (today we would use the term facies successions), and
facies”. It is understood that you have made an interpretation of various sequences and successions in particular depositional
the rocks you worked on, and that the term “fluvial facies” environments can be synthesized into models for that environ-
encompasses a constellation of features including (in the fluvial ment. Important points established in Walker (1984) include (1)
example) sharp-based fining-upward successions with lags at “the term [facies] model has a generality that goes beyond a single
their bases, thin siltstones with root traces, abundant trough and study of one formation”, and (2) the facies model is “a general
planar tabular cross bedding, and the absence of marine indica- summary of a specific sedimentary environment, written in
tors. It is normally obvious from the context whether the term terms that make the summary usable in at least four different
facies is being used in a descriptive or an interpretive sense. ways”. These ways are discussed in this present paper.
The models discussed in the first two editions of “Facies
Facies Modeling Models” (Walker, 1979, 1984) were static, inasmuch as they used
information from modern environments as seen today. The third
Although the term facies is old, the concept of facies modeling edition (Walker and James, 1992) examined models in the light of
is much younger. The term “facies model” was first used at a responses to sea level change. In the last ten years, there has been
“discussion” organized by Paul Potter at the Illinois State Geo- a large amount of new work on modern depositional environ-
logical Survey in 1958 (Potter, 1959). The purpose of the discus- ments. Also, in many cases (particularly turbidite systems) an
sion was to pool “the knowledge and experience of the group entirely new dimension has been added through 3-D seismic
concerning three topics”. These topics have a remarkably modern studies. Consequently, the first questions to be asked at the Calgary
ring to them, and could equally well have been chosen for a meeting in 2002 concerned the construction of facies models given
research conference in 2005. They involved “the existence and the size and diversity of the data base, and the appropriateness of
number of sedimentary associations; the possibility of establish- using such models in an increasingly complex world.
ing a model for each association that would emphasize the areal One of the themes of the Dallas symposium (2004) was to
distribution of lithological units within it; and the exploration of examine these questions, and perhaps to shift the emphasis onto
the spatial and sequential relations between the associations”. the individual depositional elements within environments, and
Potter reports that a facies model “was defined as the distribution the study of how these elements fit together laterally and verti-
pattern or arrangement of lithological units within any given cally. For example, submarine fans may be too complex for
association” (a “sedimentary association” being “a collection of simple models of entire fans—nevertheless, most fans contain
commonly associated sedimentary attributes”). He continued similar elements (e.g., channel fills, splays, levees, frontal lobes
with another very up-to-date concept, that “in the early stages of etc.) which can be stacked in various ways depending on relative
geological exploration, the function of the model is to improve sea level fluctuations, local tectonics, variations in sediment
prediction of the distribution of lithological types”. One of Potter’s supply and other internal and external factors. Thus the various
conclusions remains true today—“the group discussion clearly descriptive entities have been reduced to the depositional-ele-
pointed out those areas of knowledge that permit generalization, ment level, rather than the scale of the entire environment (sub-
those areas in which problems are clearly recognized and stated marine fan, delta, shoreface etc.).
but for which definitive answers are not available, and the areas The best models embody large amounts of information from
in which the problems are not as yet clearly formulated”. as many examples as possible, modern and ancient. In generaliz-
In recent correspondence, Potter (personal communication, ing this information, the resulting models can serve as reference
2005) referred back to his 1959 discussion. A “facies model” was points for interpretations of new situations and examples, and as
defined as “a commonly recurring sedimentary facies (that is a basis for making predictions from limited amounts of data in
scale independent to a large degree)”—readers should note that new situations. The predictive aspects of facies models are impor-
FACIES MODELS REVISITED 3

tant in subsurface exploration for oil, gas, and minerals, and,


increasingly, for studying and predicting the movement of ground
water through relatively unconsolidated surficial materials. In all
cases, the geometry and connectedness of the reservoirs or aqui-
fers is directly related to the original depositional environments.
Facies modeling can now be regarded as a mature science. The
basic facies, facies organizations and depositional controls of the
major environments are fairly well understood (perhaps with the
exception of submarine fans). However, depositional environ-
ments can always be subdivided (deltas, for example) and smaller-
scale models proposed for the various parts of larger-scale envi-
ronments (distributary mouth bars, interdistributary bays, pro-
grading beach ridge systems, etc.).

SCALES AND METHODS FOR DEFINING FACIES


Facies can be defined on many different scales. Whenever a
vertical stratigraphic section or core is measured, decisions have to
be made about what to include in each measurement unit. A simple
section might consist of 2 m of sandstone overlain by 2 m of shale.
The sandstone and shale units are different facies (they have a
different “aspect”), and the defining characteristic is their lithol-
ogy. In more detail, it might be decided to measure 1 m of cross-
bedded sandstone, 1 m of ripple cross-laminated sandstone, and 2
m of shale. In this case, the different sandstone facies are defined by
their sedimentary structures. It follows that the scale of subdivision
of a stratigraphic section into facies depends on:

1. the purpose of the study

2. the time available to make the measurements, and FIG. 2.—Ripple cross lamination. Red arrows show cross lamina-
tion dipping left, and blue arrows show lamination dipping
3. the abundance of descriptive features in the rocks. right. Note mudstone drapes (yellow arrows) between ripple
cross-laminated layers. Reversing flow directions with drapes
Purpose of the Study between layers suggests a tidal influence. Cretaceous, south-
ern Alberta. Core is 10 cm in diameter.
If the objective of a study is a routine description and interpre-
tation on a large scale, the scale of facies subdivision may be fairly
broad. In contrast, if the goal is a detailed analysis and interpre- an existing model, the scale of facies subdivision must be much
tation of a thin stratigraphic unit, perhaps with a view to refining more detailed. As an example, compare Figures 1, 2 and 3. They
all could be described as “ripple cross lamination”. However,
Figure 1 shows symmetrical ripple profiles with unidirectional
cross lamination, and can be interpreted as combined-flow ripples
(wave plus current). Figure 2 shows ripple cross lamination
alternating with thin mudstone drapes. The cross lamination is
directed both to the right and the left, and the assemblage of
features suggests alternating tidal currents with mud deposited
during slack tides. Figure 3 shows ripple cross lamination with-
out mudstone layers. Lamination is preserved on the ripple
crests, and sets of ripples climb on the backs of each other. This
lamination formed during active deposition from suspension,
perhaps during a waning flood stage.
These observations go beyond “ripple cross lamination”, and
involve sand/mud ratios, continuity of lamination within the
ripples and preservation (or not) of the entire ripple form. If the
cm scale purpose of the study is a detailed interpretation of sedimentary
environments, distinguishing Figures 1, 2 and 3 as different facies
FIG. 1.—Ripple cross lamination. Note symmetry of ripples (red is important. In a broader study, lumping all three Figures
arrows), but unidirectional cross lamination (to the right, together into one “ripple cross-laminated facies” might be suffi-
yellow arrows). Symmetry suggests wave action, and cross cient.
lamination suggests a superimposed unidirectional flow. These
are combined-flow ripples with relatively long periods of Time Available for the Study
mud deposition between sand emplacement. Discontinuity
of ripple layers suggests very limited sand supply. Pennsyl- With an entire day available for the description of one 18 m
vanian, Devon, England. core, it will be possible to subdivide the core into many different
4 ROGER G. WALKER

facies using varied and subtle criteria to distinguish the various


facies. In contrast, if a thick unit has to be studied over a wide area
in only two weeks of field work, the facies subdivisions will
necessarily be broader. This again is illustrated in Figures 1, 2, and
3, where more time would be necessary to distinguish and
describe the details of the ripple cross lamination, as opposed to
lumping all three figures together as “ripple cross laminated”.
The importance of time available and purpose of study are
also illustrated in Figure 4. With little time and a routine descrip-
tion on a large scale, the sandier-upward succession (red arrow)
could be described as one facies—hummocky cross-stratified
sandstones interbedded with bioturbated mudstones. If more
time were available and more detail required, the succession
could be divided into three different facies (yellow arrows), a
lower muddier facies, a central facies in which the sandstone/
shale ratio is about 1, and an upper facies consisting almost
entirely of sandstone.

Descriptive Features

FIG. 3.—Ripple cross lamination. Red arrows indicate preserva- The degree of subdivision always depends on the abundance
tion of lamination over the ripple crests, and yellow arrows of features in the rocks. A thick succession of interbedded sand-
indicate forward movement and aggradation of the ripples. stone and shales (thin-bedded turbidites) will be difficult to
This climbing-ripple cross lamination suggests rapid deposi- subdivide because of the monotonous nature of the succession. In
tion from suspension during ripple formation, with no pauses contrast, a complexly interbedded succession of mudstones and
for mud deposition. Pleistocene, southern Ontario. Coin is 2.2 sandstones with a wide variety of sedimentary structures and
cm in diameter. trace fossils, and various scales of interbedding of the lithologies

FIG. 4.—Sandier-upward succession from the Cardium Formation (Cretaceous), Blackstone River, Alberta. Person circled for scale.
Red arrow shows overall sandier-upward succession, and yellow arrows suggest how the succession could be subdivided into
a lower muddier portion, a main central portion characterized by interbedded sandstones and mudstones, and an upper
dominantly sandy portion.
FACIES MODELS REVISITED 5

(as might occur in an estuary fill), will allow detailed facies profile in Fig. 6A shows a progressive upward shift to the left,
subdivision on a small scale. Figures 1, 2 and 3 clearly present a indicating “cleaner” rocks with fewer clays—the profile is com-
wealth of descriptive features on a small scale, whereas Figure 5 monly described as cleaning-upward, or more interpretively,
presents few descriptors within the thick package of thin-bedded “sandier-upward”. The profiles in Figs. 6B and C are very similar,
sandstones and shales. and both show abrupt shifts to the left in the gamma ray profiles,
Subdivision of a stratigraphic section into facies should not rather than the progressive shift seen in Fig. 6A. The profile is
be undertaken before gaining familiarity with the complete commonly described as “blocky”. The log facies of Figs. 6A and
section. It will then be apparent how much variability there is, 6B are the same, and differ from the log facies of Fig. 6A.
and how many facies should be defined to describe the unit However, similar log facies may have very different interpreta-
adequately. Most facies studies to date have relied on qualita- tions—the blocky profiles can indicate sharp-based channel-
tively assessed combinations of characteristic sedimentary struc- filling sandstones, or sharp-based forced-regressive shorefaces.
tures and trace fossils (e.g., de Raaf et al., 1965; Williams and Without other information, preferably core control, the two pos-
Rust, 1969; Wilson, 1975; Cant and Walker, 1976; Scholle et al., sibilities cannot be distinguished.
1983; Walker, 1983). Statistical methods have also been used, The pitfalls involved in the interpretation of such log facies are
particularly where there is agreement among workers as to the highlighted by the fact that channel sandbodies and shoreface
important quantifiable descriptive parameters—for example, sandbodies commonly trend at right angles to each other.
the proportions of different types of clasts in carbonate rocks
(Imbrie and Purdy, 1962; Klovan, 1964; Harbaugh and FACIES ASSOCIATIONS
Demirmen, 1964; Harbaugh and Merriam, 1968). Statistical
methods are less suited to clastic rocks, where the most impor- In many studies, facies may have been defined in a detailed
tant descriptors (sedimentary structures and trace fossils) can- way on a small scale, with minor subtle differences between the
not easily be quantified. facies (e.g., Walker, 1983). This subdivision may result in a facies
The facies concept can be extended beyond observable rock scheme where the descriptive differences outstrip our ability to
types. Different seismic characteristics have given rise to various interpret the differences. It is therefore useful in such studies to
seismic facies (Weimer, 1989), and different well log characteris- combine closely related facies into facies associations, or “groups of
tics have also given rise to log facies (Fig. 6). The gamma ray facies genetically related to one another and which have some

FIG. 5.—Thin-bedded classical turbidites, Devonian, South Australia. Stratigraphic top to left. The succession is characterized by its
homogeneity, with almost no change in facies from bottom to top.
6 ROGER G. WALKER

05-03-50-8W5 07-20-51-11W5 05-28-09-09W4

B C

FIG. 6.—Gamma-ray logs from three locations in Alberta. A shows a “cleaner-upward” succession, whereas B and C show
examples of a blocky gamma-ray signature. Scale in meters. Interpretations: 6A is a prograding shoreface from the Cretaceous
Second White Specks, 6B is a forced regressive shoreface from the Cretaceous Cardium Formation, and 6C is a channel fill from
the Cretaceous Viking Formation.

environmental significance” (Collinson, 1969). These larger-scale 2. giant (20 m +) sets of planar tabular cross bedding (Fig. 8)
facies associations have also been termed architectural elements (eolian environments),
(Allen, 1983), denoting the building blocks of the various deposi-
tional systems. It will be suggested later that the definition of 3. thin-bedded turbidites with climbing ripples, convolute
architectural elements is fundamental to the construction of lamination, and ripped-up mud clasts (Fig. 9) (the “CCC
improved facies models in situations where the complexity of the turbidites” of Walker, 1985, interpreted as levee deposits),
system (e.g., deltas) appears to discourage the development of and
simple models.
It is now clear that some facies and architectural elements 4. cross bedding with bundles of mudstone drapes indicative
occur universally, in recent and ancient sediments, and in many of deposition in a tidally dominated environment (Fig. 10)
different basins around the world. The first universal facies (Visser, 1980).
scheme was proposed for turbidites (Mutti and Ricci Lucchi,
1972), and Miall (1977, 1985) has suggested a universal scheme for Generally, the subtle differences that enable individual facies
fluvial deposits. For example, Miall’s (1985) channel architectural to be defined may be the result of many small-scale local factors
element (CH) consists of any combination of a series of defined affecting deposition. Architectural elements are the larger-scale
lithofacies which communally have a distinctive elongate chan- components of a depositional system—they will tend to be more
nel geometry; it is part of the architecture of almost all modern general in nature, less influenced by local factors, and hence more
rivers and can be recognized in most ancient fluvial deposits. universal in their application.
Other examples of universal facies include:
FACIES SUCCESSIONS
1. sharp-based hummocky cross-stratified sandstones inter-
bedded with bioturbated mudstones (Fig. 7) (storm-domi- This term implies that certain facies properties change pro-
nated shelf deposits), gressively and systematically either vertically or horizontally.
FACIES MODELS REVISITED 7

FIG. 7.—Interbedded sandstones and mudstones, Kenilworth Formation, Book Cliffs, Utah. Red arrows show convex-upward
stratification typical of hummocky cross stratification (HCS), and black arrows show low-angle curved intersections of
stratification, also characteristic of HCS.

Examples include changing proportions of sand, giving rise to a a fall of relative sea level may force a shoreface to prograde
sandier-upward (or muddier-upward) succession (Fig. 4), chang- rapidly onto an erosion surface, placing the shoreface sharply
ing grain size giving rise to a fining-upward (or coarsening- and erosionally on shelf bioturbated mudstones. The erosion
upward) succession, or changing bed thickness giving rise to a surface represents the non-preservation of inner-shelf and shelf–
thickening-upward (or thinning-upward) succession. The im- shoreface transitional environments. Thus sharp breaks between
portance of recognizing such successions is that they place indi- facies, marked for example by channel scours, thin bioturbated
vidual facies into a context. Some individual facies, for example horizons (Glossifungites surfaces) or thin pebble lags, may signify
medium-scale cross bedding (Fig. 11), may be descriptively and fundamental changes in depositional environments, and per-
hydrodynamically identical, but may actually represent different haps the beginning of new cycles of sedimentation (de Raaf et al.,
depositional environments—medium-scale cross bedding can 1965). Many of these sharp breaks (bounding discontinuities) are
form in many settings, including meandering or braided rivers, now used to separate stratigraphic sequences and allostratigraphic
tidal inlets, a shoreface dominated by alongshore currents, or in units (discussed below).
an open marine tidal setting. The key to distinguishing the The relationships between facies within facies successions can
environments lies in the context of the facies in vertical and be shown qualitatively in facies relationship diagrams (Fig. 13; de
horizontal succession—for example, the shoreface cross bedding Raaf et al., 1965) or tabulated to show the numbers of observed
might overlie offshore mudstones and be overlain in turn by transitions. These numbers can be converted to probabilities, in a
beach and nonmarine deposits. Thus the succession contributes technique known as Markov chain analysis. This technique is not
important information that the individual facies cannot. used as commonly as it once was, but interested readers are
The relationship between depositional environments in space referred to the second edition of Facies Models (Walker, 1984;
and the resulting stratigraphic successions developed through Harper, 1984).
time was first emphasized by Johannes Walther (1894, translated
in Middleton 1973) in his Law of the Correlation of Facies. Walther FACIES MODELS
stated that:
A facies model can be defined as a general summary of a
“it is a basic statement of far reaching significance that only specific depositional environment, incorporating information
those facies and facies areas can be superimposed primarily from recent sediments and ancient rocks. Two problems are
which can be observed beside each other at the present time”. immediately apparent: scale and interpretation. The problem of
scale refers to the environment to be modeled—do we need a
Application of this law suggests that in a vertical facies model for barrier-island and lagoon systems (large scale), or
succession a gradational transition (Fig. 12) from one facies to should we be separately modeling the lagoons and the shoreface
another implies that the two facies represent environments that dunes of the barrier itself (small scale)? A simple answer
were once adjacent laterally. If the contacts between facies or suggests that both scales need modeling. The smaller scale is
facies associations are sharp and/or erosional (Fig. 12), there is no easier to define, describe, and distill, and a group of related
way of knowing whether the two vertically adjacent facies repre- small-scale models can probably be combined into a larger-
sent environments that were once laterally adjacent. For example, scale model.
8 ROGER G. WALKER

20 m

FIG. 8.—Giant (20 m) planar-tabular cross bedding from the Permian White Rim Sandstone, Canyonlands, Utah. Scale suggests an
eolian origin for the cross bedding.

The problem of interpretation concerns the integration of data such systems operate, rather than making statements only about
on ancient rocks into the facies model. For example, it is easy to each individual example (see turbidite contribution by
choose several examples of modern wave-dominated shorefaces Posamentier and Walker on this CD). These general statements
and compare their characteristics—the data set is homogeneous. are obviously more powerful than countless statements involv-
However, incorporating data from ancient sediments into this ing only individual examples. The process of extracting this
facies model involves making interpretations—we may be cor- general information is shown in Figure 14, using some modern
rect in many of the examples we choose, but some tide-dominated fans (Rhone, Amazon and Mississippi) and some ancient tur-
shorefaces may end up in our data set, which would then be less bidite examples (Wheeler Gorge, Frigg fan). Obviously, a better
homogeneous. generalization (model) would result if more examples were
In all modeling, a philosophical assumption must be made, used.
that there is system and order in Nature, and that geologists can The entire wealth of information is first distilled, boiling away
identify and agree upon a limited number of depositional envi- the local details and concentrating the important features that all
ronments and systems. In a well argued alternative view, examples share. The features that all examples share (in this case,
Anderton (1985, p. 33) suggests that “if, like me, you have a more perhaps the monotonous alternations of parallel-bedded sand-
nihilistic view of life, the universe and everything, then you stones and shales) may be relatively easy to agree upon, but in
have to admit an infinite number of environments, facies and many cases separating local detail from general principles may be
models”. more contentious. In this case, the conglomerates of Wheeler
For those of us who seek order in Nature, and who see value Gorge are quite unlike the fine-grained sheetlike turbidites of the
in building facies models, the principles, methods, and motives outer parts of Mississippi Fan.
are shown in Figure 14. In this figure, a turbidite / submarine fan Answering these questions involves a thorough knowledge
example has been used—readers should understand that in order of the literature along with extensive individual experience,
to present the basic ideas the details of submarine fan systems judgement, and discussion with other workers with different
have been very oversimplified. experiences. Models are constantly being refined as more ex-
The first assumption is that many modern submarine fan amples become available, as more distinct architectural elements
systems have been studied, and many ancient turbidite systems are recognized, and as depositional processes become better
studied and interpreted. As a result of this work, we then understood. This is shown in Figure 14 by the feedback loop from
assume that we can make some general statements about how distillation to model, from model to comparison with more local
FACIES MODELS REVISITED 9

A B C
FIG. 9.—CCC turbidites from the Cretaceous Lange Formation, offshore mid-Norway. In A, note convoluted parallel and ripple cross
lamination. In B, the ripple cross lamination becomes progressively more convoluted upward. In C, the ripples are climbing
(arrow) and convoluted. Core is 10 cm in diameter.

examples, and incorporation of those new examples into the data in 2005 may have altered the Mississippi Delta in ways that will
base. This in turn demands renewed distillation, and so on. be preserved and may be recognized in the geological record—
these effects will need to be built into future deltaic models.
CHOICE OF ENVIRONMENTS TO MODEL Similar comments concerning facies complexity and the impor-
tance of rare catastrophic events can be made about submarine
Facies models have traditionally been formulated for deposi- fans and fluvial systems.
tional systems that form obvious geographical entities—for ex-
ample, meandering rivers or deltas. In many environments, and Depositional Elements Rather Than Environments
deltas are a good example, it is clear that many different deposi-
tional processes combine to give many different depositional Instead of modeling obvious geographical entities, it may be
results. Thus deltas may be wave dominated, river dominated, or preferable to model discrete depositional elements. These ele-
tide dominated, all of which have very different geometries and ments may be found in several geographical settings. An obvious
sandbody distributions (Bhattacharya, this CD). Yet they all example is the shoreface depositional element. It is controlled by
conform to one of the classic definitions of a delta, as a “river-fed alongshore sediment supply under the influence of shoaling waves.
depositional system that results in an irregular progradation of Variability of process, for example the relative importance of fair-
the shoreline” (Scott and Fisher, 1969). weather and storm processes, combined with the rate and caliber
The results of nearly fifty years of research (since publication of sediment supply, combine to give rise to a closely related set of
of “Recent Sediments, Northwest Gulf of Mexico”; Shepard et al., depositional products. If the immediate snapshot of the shoreface
1960) have suggested that deltas may be too big and too complex at one time is combined with evolution of the depositional system
for the formulation of good facies models. Variability within through time, emphasizing relative sea-level fluctuations, the depo-
river-influenced deltas such as the Mississippi is enormous, sitional products may be gradationally based and sandier upward
embracing many smaller environments that may deserve their (a normal prograding shoreface), or sharp based and sandy through-
own models (distributary mouth bars, bays, distributary chan- out (the result of progradation during relative sea level fall—a
nels and levees, crevasses splays, beaches and barriers, lagoons, forced regressive sequence). Nevertheless, the shoreface is a rela-
and offshore shoals, among others). It should also be emphasized tively easily defined and easily understood depositional element
that the basis of many facies models consists of the work done (compared with the complexity of a delta). The shoreface element
during long periods of “normal” conditions. Hurricane Katrina occurs in several geographic environments, and contributes to
10 ROGER G. WALKER

FIG. 10.—Set of cross bedding 20 cm thick from Cretaceous sandstones of Leighton Buzzard, England. Note mudstone drapes on the
foresets (yellow arrows), and at least one paired set of drapes (red arrow) indicative of deposition in a tidal setting. The mudstone
drapes form during slack tides.

several geographically based traditional facies models—for ex- tion by Boyd et al. on this CD), where one of the most obvious
ample, wave-dominated deltas, prograding strandplains, barrier parts of the system involves the foreshore and barrier superstruc-
islands and transgressive shoreline systems. Other examples are ture (Fig. 15). However, Rampino and Sanders (1980) have shown
discussed in the individual contributions on this CD. by detailed coring studies (Fig. 16) that during transgression the
Another excellent example of depositional elements within a sand from the foreshore and barrier superstructure is moved (1)
specific depositional environment has been presented by Miall seaward by storm waves to form new nearshore sand ridges (Fig.
(1977, 1985). In this fluvial example, a set of architectural elements 16, red arrow), and (2) landward into the lagoon as washover
has been defined, using capital letters to designate grain sizes (G facies (Fig. 16, blue arrow). In a transgressive setting, the barrier
for gravel, S for sand, etc.). Lower-case letters were used to itself is not preserved—the resulting sedimentary record is shown
indicate sedimentary features (f for flat bedded, for example), in the black rectangle of Fig. 16, and consists of thin lagoonal and
resulting architectural element designations such as Gf. Miall’s washover deposits, sharply overlain by thin nearshore sand
catalog of elements for fluvial systems is very useful, particularly ridges. Any attempt to model a geographically defined barrier
for workers new to fluvial systems seeking guidance in what to island and lagoonal system is bound to encounter severe prob-
look for and what to describe. More experienced workers will be lems of facies preservation. It will be more fruitful to identify the
sensitive to the possibility that there may be elements that do not various depositional elements of the system, to study the pro-
fall easily into Miall’s catalog. cesses that control them today, and the ways in which they will
(or will not) be preserved in the geological record. Our model, for
Elements Remain Constant— purposes of comparison and prediction (see below), may end up
Geographic Environments Change Through Time as a “transgressive lagoon-washover” model (Fig. 16), rather than
a “barrier-island” model exemplified by Figure 15.
The shoreface depositional element discussed above remains
relatively constant through time. It progrades given sufficient THE USES OF FACIES MODELS
sediment supply, with details of the progradation being controlled
by wave and tidal processes, and changing relative sea level. There has been little or no discussion in the literature of the
On a larger scale, geographically defined environments may original four uses of facies models proposed by Walker (1979,
change dramatically through time, such that very few parts of 1984; Walker and James, 1992). The generality embodied in a
today’s snapshot may be preserved in the geological record. model, as opposed to a summary of one particular example,
Barrier islands form an excellent example (Fig. 15; see contribu- enables the model to assume four main functions (Fig. 14):
FACIES MODELS REVISITED 11

FIG. 11.—Trough cross bedding, seen more or less parallel to flow direction (to the left), from the Cretaceous rocks in, Berry Gulch,
Colorado. 15 cm scale circled.

1. a norm, for purposes of comparison a stack of two 3-m-thick sandbodies? These are questions and
ideas that are possible only if the new example is compared with
2. a framework and guide for future observations a norm. A fourth question also arises: is the comparison valid, or
is an apple being compared with a norm for oranges? In the case
3. a predictor in new situations, and of Allen’s fluvial successions, the norm has probably been con-
structed from very homogeneous data. Comparisons with other
4. a basis for interpretation Devonian examples may be good, but comparisons with (say)
Cretaceous rivers from very high-accommodation settings may
The Model As a Norm be less useful.

Figure 17 shows a simple 2-D model of a fluvial fining- The Model As a Framework for Observations
upward succession. It was derived from data published by
Allen (1970) on more than one hundred examples of Devo- A good model summarizes all of the important descriptive
nian fluvial successions in Britain, and redrawn to scale in features of a particular system. For example, in Figure 17 the
Figure 17. It is characterized by roughly equal proportions of fluvial fining-upward sequence contains a basal lag overlain
point-bar and vertical-accretion facies, both about 3 m thick, by various cross-bedded and parallel-laminated facies. These
and the entire fining-upward sequence can be considered as in turn are overlain by ripple cross-laminated facies. The
a norm. vertical-accretion facies may contain root traces, desiccation
Let us then assume that during field work a new succession is cracks, and caliche nodules. These are the basic descriptors of
found with a 6-m thick point-bar succession and 3 m of vertical this particular model, and they act as a guide for making
accretion facies. By itself, this succession may be difficult to observations in new examples—is the succession the same; are
interpret, but by comparison with the norm (Fig. 17) it is imme- the proportions of facies the same; are any distinctive features
diately clear that the new point-bar succession is twice as thick as absent; or are there new features that are not included in the
the norm. This comparison opens new lines of thought and current model? Miall’s (1977, 1985) fluvial depositional ele-
interpretation—was the river unusually deep; was the rate of ments (Gf, etc.) are also excellent examples of a framework for
subsidence unusually high; does the sandbody actually consist of future observations.
12 ROGER G. WALKER

Sharp and/or erosive facies boundaries

Gradational facies boundaries


within successions

FIG. 12.—Cretaceous Mountain Park Formation, Alberta. Yellow arrows show three sandier-upward successions. Within these
successions, all of the facies boundaries are gradational. At the tops of the successions, contacts are sharp (yellow dotted lines).
At the horizon of the uppermost sandstone, the dotted lines show a sharp base and a sharp and erosional top (seen at right).

The Model As a Predictor incorrectly interpreted as distal basin-plain facies, the wrong
model might be chosen for prediction. It follows that the second
This is without question the most important function of any important step involves selection of an appropriate model. As
facies model. The basic idea is very simple: given one new piece another example, let us assume that our new piece of information
of information, it may be possible (1) to assign that information to consists of the lag and cross-bedded sandstones at the base of the
a particular model, and therefore (2) use the model to predict the succession in Figure 17. The appropriate model would be a
rest of the system. meandering-river model—an inappropriate model would be a
As an example, some thin-bedded turbidites are shown in braided-river model or a tidal-inlet model.
Figure 9. The sandstones are characterized by climbing ripples In many cases, it will be possible to test the predictions made
and convolute lamination, features that are more abundant in from the one new piece of information, perhaps by examining all
thin-bedded levee turbidites than in thin-bedded basin-plain of the nearby outcrops. But if thin-bedded turbidites similar to
turbidites (Walker, 1985). We may therefore make a preliminary those in Figure 9 are also characteristic of one core from a Miocene
interpretation, assigning the beds in Figure 9 to the levee of a submarine fan in offshore West Africa, testing the prediction of a
deep-sea channel. To make further predictions, we select a model nearby channel sandbody may involve millions of dollars of
for deep-sea channels (see Posamentier and Walker on this CD). drilling costs. Clearly, the new data must be interpreted as
The one new piece of information (thin-bedded levee facies), plus carefully and accurately as possible, and the most appropriate
the appropriate model, suggests a system involving back-of- model used for prediction.
levee facies, channel-margin facies, channel-fill facies (perhaps
coarser sandstones interbedded with mudstones), and possible The Model As a Basis for Hydrodynamic Interpretation
channel “point-bar” facies with morphological scars as the point
bar has shifted in position as the channel migrated (Kolla et al., This use of facies models was originally prompted by the idea
2001; Posamentier and Walker on this CD). that one individual turbidite may be difficult to interpret, whereas
There are two equally important steps in using models as many turbidites combined into a model (the Bouma sequence,
predictors. The first involves the correct interpretation of the new 1962) would give a more consistent and general basis for inter-
piece(s) of information. If thin-bedded levee turbidites were preting depositional processes (Harms and Fahnestock, 1965;
FACIES MODELS REVISITED 13

cross-stratified sandstones interbedded with bioturbated


mudstones (Fig. 7), b) thin-bedded turbidites with climbing
ripples, convolute lamination, and ripped-up mud clasts
(Fig. 9), c) sets of cross stratification thicker than 10 m (Fig.
8), that may be either planar-tabular or trough shaped (the
latter commonly with very long toesets), and d) cross-
bedded sandstones with paired mudstone drapes on the
foresets (Fig. 10). There are more and more examples of
universal facies being recognized, and these will form the
basic building blocks of the sedimentary record.

2. Modern processes, varying rates of sediment supply, and


fluctuations in relative sea level combine to form distinctive
depositional elements in recent sediments. Some of these
elements are also universal, and occur in many different
places around the world today. Examples include a)
shorefaces, b) eolian desert dune complexes, c) tidal inlet–
tidal delta systems, and d) lagoonal and barrier washover
systems. Many other examples could easily be added to this
list, but perhaps with the exception of desert dunes the
depositional elements tend to be smaller rather than larger
in scale, and homogeneous in character.

Granted these two “truths”, the future of modeling may lie in


FIG. 13.—Facies-relationship diagram from the Carboniferous
refining and agreeing upon the commonly occurring universal
Westward Ho! Formation, north Devon, England. Facies have
architectural elements in the geological record (a process akin to
been given descriptive names, and arrows show sharp and
distillation in Fig. 14). At the same time, there may be more
gradational facies contacts, and numbers indicate the occur-
refining and agreeing upon the combinations of processes that
rence of each transition. From de Raaf et al. (1965).

Walker, 1965). This interpretive usage of models is probably more


effective for small-scale models (the Bouma sequence for turbid-
ites), and less effective for large-scale systems (e.g., deltas). Also,
individual turbidity-current depositional processes are difficult Distill local examples

to observe in modern oceans, whereas work over the last 20 years


in nonmarine and shallow-marine depositional environments
has added enormously to our understanding of depositional Make a model
processes. Models have not been used extensively as a basis for
hydrodynamic interpretation, and this is probably their least
useful aspect.
Comparison
FUTURE APPROACHES TO MODELING ▼

It was pointed out above that our knowledge of modern


environments has expanded tremendously in the last twenty-five
years (since the publication of the first edition of Facies Models).
Because of this, many modern environments may be perceived as
so complex that simple models and distillations may be impos- ▼
sible, inappropriate, or both. In the absence of models, however,
there will be no norms, and no bases for making predictions. In
the hope that modeling of any sort is preferable to anarchy, I
emphasize two basic “truths” from which we might proceed. Prediction
1. Individual, small-scale facies can be identified in ancient
rocks, as shown in Figures 1, 2 and 3. Associations of facies
that commonly occur laterally and vertically adjacent to each
other can be combined into facies associations, and these
associations (or architectural elements) form the basis for a
descriptive subdivision of the stratigraphic record. The na- FIG. 14.—Facies modeling, from Walker and James (1992). Note
ture of the bounding surfaces between associations/ele- the relationship between individual examples and their
ments is also important—they may be gradational, or sharp distillation into a general model. Note how new examples
and/or erosional (Fig. 12). It is becoming apparent that some can be compared with the model (the “norm”), and then
facies, facies associations, and architectural elements are incorporated into the general data base (feedback). The
universal (i.e., they occur in many different places, in all parts model also serves as a guide for making observations and as
of the geological record). Examples include a) hummocky a predictor in new situations.
14 ROGER G. WALKER

FIG. 15.—Block diagram of a barrier-island and lagoonal depositional environment, from Reinson (1992). Note that this is a “snapshot
in time” and that the various environments shown in the diagram may not be preserved if the barrier progrades or is transgressed.
For the transgressive setting, see Figure 16.

Washover

Superstructure
Foreshore
Nearshore
ridges

FIG. 16.—Cross section of Cedar Beach, Long Island, New York. Barrier–lagoon system rests disconformably on Pleistocene diamictite
(blue, conglomerate–breccia symbols). The barrier cuts down into the Pleistocene at a tidal inlet. Note the age dates of lagoonal
deposits, 7815 and 7130 years BP south of the barrier, 5055 years BP at Cedar beach, and 1015 and 300 years BP at the northern
edge of the lagoon. The sand of the barrier superstructure is not preserved during transgression—it is washed into the lagoon
(dark red) during storms (blue arrow) and is also moved offshore to form small ridges (red arrow—southern end of cross section).
The record of barrier transgression is shown in the black rectangle, and consists of a thin sheet of lagoonal and washover facies
(blue, dark red, seaward of barrier) overlain disconformably by modern storm sands. The two black arrows at the right show two
erosion surfaces, one separating Pleistocene from Holocene sediments and one separating 7000+ year-old lagoonal and washover
facies from offshore sands forming today. From Rampino and Sanders (1970).
FACIES MODELS REVISITED 15

7 nearshore ridges, lagoon, and washover. Each association will


have distinctive lithologies, sedimentary structures, and trace-
fossil assemblages. However, if the situation shown in Figure 15
CALICHE were to be preserved in the geological record, three of the associa-

MUDSTONES 2.86 m
NODULES tions would be closely related (the lagoon, washover, and near-

VERTICAL ACCRETION
6 shore ridges shown in the black rectangle in Fig. 16), and two
DESSICATION would be missing (barrier superstructure and foreshore). Signifi-
cantly, the lagoon and washover facies would lie on a major

FLOODPLAIN
CRACKS
transgressive surface of erosion, and the nearshore ridges would
be separated from the lagoon and washover facies by another
ROOT transgressive surface of erosion (Fig. 16). Thus the geological
5 TRACES interpretation would be based on (1) defining the facies associa-
tions present, (2) making a preliminary interpretation of the
associations, and (3) defining the relationships of the facies asso-
ciations and their bounding surfaces. Item 3 in this list is the crux
of the overall interpretation, because it involves building the
4 facies associations into a three-dimensional structure, indepen-
ALTERNATING dently of any preconceived facies model.
SSTS. AND MSTS.
1.00 m BUILDING MODELS
3 The ideas presented above suggest that we are moving away
RIPPLE from interpretation by reference to existing models. We are better
LATERAL ACCRETION
CROSS-LAM. able to recognize facies and architectural elements. We have more
0.95 m data expressing the complexity of modern environments. It there-
fore becomes less and less appropriate to use simple models to
2
POINT BAR

interpret complex geological situations.


The solution is not to abandon models, which would result in
anarchy. The solution is to build your own interpretations, using
TROUGH, P.T. the following stages.
CROSS BED
1 1.71 m 1. Recognize and define facies, facies associations, and archi-
tectural elements in the example you are studying. Some of
the elements may be universal, and some may be local to
your particular example.
LAG. 0.32 m
0 2. Carefully fit the elements into their 3-D framework. Which
METERS ones occur together, and which are never found together?
Define the surfaces that separate the elements.
FIG. 17.—Fining-upward meandering-fluvial succession, com-
plied by Walker (1979) from data published by Allen (1970). 3. Attempt a preliminary interpretation of those elements that
The average thicknesses shown for each facies are also com- allow it. Some will probably present features that have been
piled from Allen’s data. well studied and have agreed-upon interpretations (e.g.,
hummocky cross stratification), but other may be enigmatic
(e.g., thick structureless sandstones).
form distinctive depositional systems in recent sediments. It
follows that the combination of architectural elements in ancient 4. From whatever interpretation is possible, refer to the clos-
sediments, and depositional systems in recent sediments, will est existing model (the model as a norm). How does the
form the true basis for defining the building blocks of the sedi- distribution of depositional elements in the model conform
mentary record. As an example, I again refer to the shoreface, to the distribution defined in your example? Lagoonal and
where the architectural element is well defined in the geological washover elements might suggest reference to a barrier-
record and where there is a large body of data from recent island model, but the association of lagoonal, washover,
sediments. Questions remain, particularly concerning the con- and nearshore-ridge elements defines one part of a dy-
cept and definition of fair-weather wave base, and the preserv- namic barrier system rather than a barrier–lagoon system
ability of many of the features seen in modern shorefaces. Never- shown as a block diagram (e.g., Reinson, 1992, his fig. 3).
theless, the shoreface is an important example of a basic architec- This approach might be even more important in submarine
tural and depositional element, particularly in its role in defining fan systems, where there has been considerable work on
transgressions and regressions. depositional elements but no simple fan models since the
In many ancient examples from the geological record, there 1970s and early 1980s.
may be no easy and direct comparison with existing facies mod-
els. It therefore becomes even more important to recognize indi- Models and Interpretations
vidual facies and facies associations (architectural elements), and
to determine which elements commonly occur together and The three-dimensional relationships of architectural ele-
which never occur together. In Figure 15, at least five associations ments that emerge from a new study of the hypothetical Beau-
could be recognized: barrier superstructure, barrier shoreface, fort Sea Formation add up to an interpretation of that formation,
16 ROGER G. WALKER

and NOT to a model for the Beaufort Sea Formation. There are HARPER, C.W., JR., 1984, Improved methods of facies sequence analysis, in
many papers in the literature that have titles along the lines of Walker, R.G., ed., Facies Models, Second Edition: Geological Associa-
“The Beaufort Sea Formation—a model for deposition in shal- tion of Canada, Reprint Series 1, p. 11–13.
low marine environments”. Presumably the idea is that if a IMBRIE, J., AND PURDY, E.G., 1962, Classification of modern Bahamian
model is presented, a routine description of the Beaufort Sea carbonate sediments, in Ham, W.E., ed., Classification of Carbonate
Formation may sound more interesting, and hence attract atten- Rocks: American Association of Petroleum Geologists, Memoir 1, p.
tion. 253–279.
One formation may provide a superb case history, but one KLOVAN, J.E., 1964, Facies analysis of the Redwater reef complex, Alberta,
formation does not make a model. The theme of facies modeling Canada: Bulletin of Canadian Petroleum Geology, v. 12, p. 1–100.
is one of distilling many examples in the search for generality. KOLLA, V., BOURGES, P., URRUTY, J.M., AND SAFA, P., 2001, Evolution of deep-
Small is beautiful because small is usually more homogeneous. I water Tertiary sinuous channels off shore Angola (west Africa) and
therefore suggest that one important approach in the future will implications for reservoir architecture. American Association of Pe-
be to define the pieces (modern and ancient) and to define the troleum Geologists, Bulletin, v. 85, p. 1373–1405.
relationships between the pieces. This is part of the distillation MIALL, A.D., 1977, A review of the braided river depositional environ-
process, but it begins with the pieces rather than with an initial ment: Earth-Science Reviews, v. 13, p. 1–62.
assumption of a geographically defined environment (delta, MIALL, A.D., 1985, Architectural element analysis: a new method of facies
whatever). Progress will truly be made when the geologically analysis applied to fluvial deposits: Earth-Science Reviews, v. 22, p.
defined pieces (ideally, the universally accepted architectural 261–308.
elements) closely agree with the depositional elements defined MIDDLETON, G.V., 1973, Johannes Walther’s law of the correlation of facies:
by modern processes. Geological Society of America, Bulletin, v. 84, p. 979–988.
MIDDLETON, G.V., 1978, Facies, in Fairbridge, R.W., and Bourgeois, J., eds.,
ACKNOWLEDGMENTS Encyclopedia of Sedimentology: Stroudsburg, Pennsylvania, Dowden,
Hutchinson & Ross, p. 323–325.
I thank Brian A. Zaitlin and Henry Posamentier for their MUTTI, E., AND RICCI LUCCHI, F., 1972, Le torbiditi dell’Appennino
comments on the manuscript. I also thank the Natural Sciences settentrionale: introduzione all’analisi de facies: Societá Geologica
and Engineering Research Council of Canada for their support of Italiana, Memorie, v. 11, p. 161–199. English translation by T.H.
my research. Nilsen, 1978, International Geology Review, v. 20, p. 125–166.
POTTER, P.E., 1959, Facies models conference: Science, v. 129, p. 1272–1273.
REFERENCES RAMPINO, M.R., AND SANDERS, J.E., 1980, Holocene transgression in south-
central Long Island, New York: Journal of Sedimentary Petrology, v.
ALLEN, J.R.L., 1970, Studies in fluviatile sedimentation: a comparison of 50, p. 1053–1079.
fining-upward cyclothems, with special reference to coarse member REINSON, G.E., 1992, Transgressive barrier island and estuarine systems, in
composition and interpretation: Journal of Sedimentary Petrology, v. Walker, R.G., and James, N.P., eds., Facies Models: Geological Asso-
40, p. 298–323. ciation of Canada, p. 179–194.
ALLEN, J.R.L., 1983, Studies in fluviatile sedimentation: bar complexes and SCHOLLE, P.A., BEBOUT, D.G., AND MOORE, C.H., 1983, Carbonate Deposi-
sandstone sheets (low sinuosity braided streams) in the Brownstones tional Environments: American Association of Petroleum Geologists,
(L. Devonian), Welsh Borders: Sedimentology, v. 33, p. 237–293. Memoir 33, 708 p.
ANDERTON, R., 1985, Clastic facies models and facies analysis, in Brenchley, SCOTT, A.J., AND FISHER, W.L., 1969, Delta systems and deltaic deposition,
P.J., and Williams, B.J.P., eds., Sedimentology: Recent Developments in Fisher, W.L., Brown, L.F., Scott, A.J., and McGowen, J.H., collo-
and Applied Aspects: Oxford, U.K., Blackwell Scientific Publications, quium leaders, Delta Systems in the Exploration for Oil and Gas:
p. 31–47. Texas Bureau of Economic Geology, Research Colloquium, various
BOUMA, A.H., 1962, Sedimentology of Some Flysch Deposits: Amsterdam, pagination.
Elsevier, 169 p. SHEPARD, F.P., PHLEGER, F.B., AND VAN ANDEL, T.H., 1960, Recent Sediments,
CANT, D.J., AND WALKER, R.G., 1976, Development of a braided fluvial Northwest Gulf of Mexico: American Association of Petroleum Ge-
facies model for the Devonian Battery point Formation, Quebec: ologists, 394 p.
Canadian Journal of Earth Sciences, v. 13, p. 102–119. TEICHERT, C., 1958, Concepts of facies: American Association of Petroleum
COLLINSON, J.D., 1969, The sedimentology of the Grindslow Shales and the Geologists, Bulletin, v. 42, p. 2718–2744.
Kinderscout Grit: a deltaic complex in the Namurian of northern VISSER, M.J., 1980, Neap–spring cycles reflected in Holocene sub-tidal
England: Journal of Sedimentary Petrology, v. 39, p. 194-221. large-scale bedform deposits: a preliminary note: Geology, v. 8, p.
DE RAAF, J.F.M., READING, H.G., AND WALKER, R.G., 1965, Cyclic sedimenta- 543–546,
tion in the lower Westphalian of North Devon, England: Sedimentol- WALKER, R.G., 1965, The origin and significance of the internal sedimen-
ogy, v. 4, p. 1–52. tary structures of turbidites. Yorkshire Geological Society, Proceed-
GRESSLY, A., 1838, Observations géologiques sur le Jura Soleurois: Neue ings, v. 35, p. 1–32.
Denksch. allg. schweiz., Ges. ges. Naturw., v. 2, p. 1–112. WALKER, R.G., 1975, Generalized facies models for resedimented con-
HARBAUGH, J.W., AND DEMIRMEN, F., 1964, Application of factor analysis to glomerates of turbidite association: Geological Society of America,
petrologic variations of Americus Limestone (Lower Permian), Kan- Bulletin, v. 86, p. 737–748.
sas and Oklahoma: Kansas Geological Survey, Special Distribution WALKER, R.G., ed., 1979, Facies Models: Geological Association of Canada,
Publication 15, 50 p. 211 p.
HARBAUGH, J.W., AND MERRIAM, D.F., 1968, Computer Applications in WALKER, R.G., 1983, Cardium Formation 3. Sedimentology and stratigra-
Stratigraphic Analysis: New York, Wiley, 262 p. phy in the Garrington–Caroline area: Bulletin of Canadian Petroleum
HARMS, J.C., AND FAHNESTOCK, R.K., 1965, Stratification, bed forms and flow Geology, v. 31, p. 213–230.
phenomena (with example from the Rio Grande), in Middleton, G.V., WALKER, R.G., ed., 1984, Facies Models, Second Edition: Geological Asso-
ed., Primary Sedimentary Structures and Their Hydrodynamic In- ciation of Canada, 317 p.
terpretation: Society of Economic Paleontologists and Mineralogists, WALKER, R.G., 1985, Mudstones and thin-bedded turbidites associated
Special Publication 12, p. 84–115. with the Upper Cretaceous Wheeler Gorge conglomerates, Califor-
FACIES MODELS REVISITED 17

nia: a possible channel–levee complex: Journal of Sedimentary


Petrology, v. 55, p. 279–290.
WALKER, R.G., AND JAMES, N.P., eds., 1992, Facies Models: Response to Sea
Level Change: Geological Association of Canada, 409 p.
WEIMER, P., 1989, Sequence stratigraphy of the Mississippi Fan (Plio-
Pleistocene), Gulf of Mexico: Geo-Marine Letters, v. 9, p. 185–272.
WILLIAMS, P.F., AND RUST, B.R., 1969, The sedimentology of a braided river:
Journal of Sedimentary Petrology, v. 39, p. 646–679.
WILSON, J.E., 1975, Carbonate Facies in Geologic History: New York,
Springer-Verlag, 471 p.
18 ROGER G. WALKER
EOLIAN FACIES MODELS 19

EOLIAN FACIES MODELS

NIGEL P. MOUNTNEY
Earth Sciences and Geography, Keele University, Keele, Staffordshire, ST5 5BG, UK
e-mail: n.p.mountney@keele.ac.uk

ABSTRACT: Although eolian facies models have been developed since the 1970s, only recently have they become sufficiently sophisticated
to enable the effects of external climatic and tectonic controls to be expressed in terms of resultant facies architecture. By using a joint
conceptual and process-based approach, the response of eolian systems to changes in controlling parameters such as sediment supply,
sediment availability, water table, and wind regime is now well understood. Dynamic facies models are able to account for spatial and
temporal variations in these controlling parameters and predict likely stratigraphic responses. Large-scale, quantitative stratigraphic data
sets from outcrop are being applied to unequivocally demonstrate relationships between preserved eolian architecture and original
bedform morphology and migratory behavior.
In dry eolian systems, the key to developing predictive facies models has been an appreciation of the paleoenvironmental significance
of the 3D geometry and hierarchical nature of bounding surfaces, which has enabled the products of external (allocyclic) controls such
as climate change be discerned from the complex mechanics of intrinsic (autocyclic) bedform migratory behavior. In wet eolian systems,
subtle variations in interdune architecture provide the basis for a spectrum of predictive facies models that explain preserved eolian
architecture in terms of interactions between water-table level, sediment availability, dune size, and dune migration rate, parameters
which in turn are a function of sediment distribution pathways, climate, and basin setting.
The development of eolian facies models is important for understanding the likely response of desert systems to climatic and
environmental change. Additionally, predictive models remain important for hydrocarbon exploration, particularly in mature
provinces, where good well control allows the employment of sophisticated models in the search for small plays based on subtle
stratigraphic traps.

INTRODUCTION HISTORY OF THE DEVELOPMENT


OF EOLIAN FACIES MODELS
Although eolian processes operate in a variety of deposi-
tional settings (e.g., beaches, fluvial and glacial outwash plains, The development of eolian facies models has been ongoing
and volcanic regions) and under the influence of various since the 1960s, and today’s sophisticated models incorporate
climates, their occurrence is most closely associated with hot- ideas that have evolved since that time (Table 1).
and cold-climate arid systems, and this chapter therefore fo-
cuses predominantly on the development of facies models for The Early Years
desert eolian systems. Recognition of ancient eolian deposits
can sometimes be straightforward. For example, very large- The first in-depth study of eolian sediments was undertaken
scale sets of cross bedding exposed across much of the south- by Bagnold (1941), who investigated the mechanics of sediment
western United States have long been ascribed an eolian dune entrainment, transport, and deposition. McKee (1966) and Th-
origin (Huntington, 1907, referenced in McKee, 1979). How- ompson (1969) conducted detailed early studies of modern and
ever, in many cases recognition is problematic, principally ancient eolian strata, respectively, whilst Stokes (1968) pro-
because it can often be difficult to differentiate strata of eolian posed that many eolian systems underwent periodic deflation
origin from the deposits of other environments, notably those down to the level of the water table and that extensive bounding
of sandy fluvial and shallow marine origin. Furthermore, surfaces were likely to represent regional deflation surfaces,
building a detailed representation (model) of the architecture later to become known as “Stokes surfaces”. Although these
of an eolian succession and devising a paleoenvironmental works illustrated the 3D complexity of eolian strata and pro-
reconstruction is also problematic because (1) although eolian vided the first rudimentary models that related modern dunes
dune systems are typically dominated by only three or four to preserved sets of cross bedding, it was to be a further decade
facies types, variations in the geometric arrangement of these before unifying models were developed. Throughout the 1970s
facies can lead to radically different models, (2) although other there was a growing need for predictive eolian facies models
facies make up only a small proportion of most successions, because a number of significant hydrocarbon provinces utilized
their interpretation is critical to determining the mechanisms reserves with eolian reservoirs. In particular, the Permian
that controlled accumulation, (3) erosion is an inherent part of Rotliegend Group of the Southern North Sea was a major source
eolian bedform migration, and preserved successions are al- of gas (Glennie, 1972), and a better understanding of reservoir
ways highly fragmentary, and (4) most eolian successions are architecture was crucial to the successful exploitation of these
characterized by marked lateral variations in bed thickness reserves.
and a low degree of lateral continuity, such that traditional 1D At White Sands, New Mexico, McKee and Moiola (1975)
sedimentary logs are not particularly useful for interpretation. demonstrated that large eolian bedforms climbed downwind as
For these reasons coherent eolian facies models were not they migrated, such that the basal parts of eolian dunes became
developed until the late 1970s, somewhat later than for most preserved because they were overridden by succeeding dunes.
other environments. Brookfield (1977) introduced a model that explained the origin of

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 19–83.
20 NIGEL P. MOUNTNEY

TABLE 1.—Summary of notable studies of modern and ancient eolian systems and important conceptual and modeling studies.
The modern and ancient eolian systems summarized are representative examples, and the list is not comprehensive.
Classic studies that are recommended as an initial source of further reading are highlighted in bold.

Studies of modern and recent eolian systems


Eolian processes
General studies Bagnold (1941)
Cooke et al. (1993)
Glennie (1970)
Kocurek (1996)
Lancaster (1995)
Erg dynamics and behavior Wilson (1971, 1973)
Eolian bedforms
General studies Glennie (1970)
McKee (1979)
Ripples Anderson (1987)
Ellwood et al. (1975)
Fryberger and Schenk (1981)
Fryberger et al. (1992)
Sharp (1963)
Dunes and draa Lancaster (1988, 1998)
McKee (1979)
Sweet (1992)
Sweet and Kocurek (1990)
Werner and Kocurek (1999)
Wilson (1971)
Structures and stratification
Wind ripple, grainflow and grainfall Anderson (1988)
strata Hunter (1977, 1981)
Fryberger and Schenk (1981, 1988)
Kocurek and Dott (1981)
Rubin (1987a)
Adhesion strata Kocurek and Fielder (1982)
Olsen et al. (1989)
Soft sediment deformation Doe and Dott (1980)
Horowitz (1982)
McKee et al. (1971)
Cross bedding and bounding Bristow et al. (1996, 2000, 2004)
surfaces McKee (1966)
Remote sensing Breed and Grow (1979)
Breed et al. (1979, 1987)
Inland eolian dune–interdune systems
Ackchar Erg Mauritania Kocurek et al. (1991a)
Al Liwa Sand Sea Abu Dhabi Bristow et al. (1996)
Duero Basin Dune Field Central Spain García-Hidalgo (2002)
Gran Desierto Sand Sea Mexico Lancaster et al. (1987)
Great Sand Dunes Colorado Andrews (1981)
Fryberger (1990a)
Kelso Dune Field Mojave Desert Kocurek and Lancaster (1999)
Nebraska Sand Hills Nebraska Ahlbrandt and Fryberger (1980)
Loope et al. (1995)
Namib Sand Sea Western Namibia Lancaster (1985)
Lancaster and Teller (1988)
Skeleton Coast Dune Field Northern Namibia Stanistreet and Stollhofen (2002)
White Sands New Mexico Fryberger et al. (1988)
McKee (1966)
McKee and Moiola (1975)
Simpson and Loope (1985)
Coastal eolian dune–interdune systems
Guerrero Negro Baja California, Mexico Fryberger et al. (1990)
Jafurah Saudi Arabia Fryberger (1990c)
Fryberger et al. (1984, 1988)
Oregon Coastal Dunes Oregon Hunter and Richmond (1988)
Hunter et al. (1983)
Padre Island Texas Hummel and Kocurek (1984)
Kocurek et al. (1992)
Schenk (1990)

geometrically complex sets of strata in terms of a hierarchy of an eolian origin. Other important developments in the 1970s
bounding surfaces that originated in response to interdune mi- included the comprehensive studies of the sedimentary features
gration in front of large bedforms, the migration of smaller of modern sand seas by Glennie (1970), and studies by Wilson
superimposed bedforms over larger bedforms, and the periodic (1971, 1972, 1973) on sand-sea dynamics and their response to
reactivation of dune lee slopes in response to changes in wind external controls. The decade culminated in the publication of A
direction. Hunter (1977) proposed criteria for the identification of Study of Global Sand Seas (McKee, 1979), which remains an invalu-
small-scale eolian sedimentary structures, thus enabling many able reference. Technological advances in the late 1970s provided
hitherto ambiguous successions to be ascribed unequivocally to a new view of eolian systems through the acquisition of remotely
EOLIAN FACIES MODELS 21

TABLE 1 (continued).—

Studies of modern and recent eolian systems (continued)


Sand sheets
General Kocurek and Nielson (1986)
Koster (1988)
Algodones California Nielson and Kocurek (1986)
Askja Central and NE Iceland Mountney and Russell (2004)
Dhahran Saudi Arabia Fryberger et al. (1983)
Salima Eastern Sahara, NE Africa Breed et al. (1987)
Maxwell and Haynes (1989)
Kuwait Desert Arabian Gulf Khalaf (1989)
Khalaf et al. (1984)
Namib Sand Sea Western Namibia Lancaster (1985)
Ténéré Desert Niger Warren (1971)
Sabkhas
Bahrain Coast Bahrain Doorncamp et al. (1980)
Dhahran Saudi Arabia Fryberger et al. (1983)
Trucial Coast Arabian Gulf Evans et al. (1964)
Pursar and Evans (1973)

Studies of ancient eolian systems


Dry eolian systems
Botucatu Sandstone Formation Cretaceous, Brazil Scherer (2001)
Etjo Sandstone Formation Cretaceous, NW Namibia Mountney and Howell (2000)
Mountney et al. (1999a,b)
Jerram et al. (1999)
Navajo Sandstone Jurassic, SW Utah Herries (1993)
Middleton and Blakey (1993)
Page Sandstone Jurassic, Utah and N Arizona Blakey et al. (1996)
Havholm and Kocurek (1994)
Havholm et al. (1993)
Kocurek et al. (1991b)
Wingate Sandstone Jurassic, SE Utah Clemmensen and Blakey (1989)
Clemmensen et al. (1989, 1994)
Wet eolian systems
Dala Sandstone Proterozoic, Sweden Pulvertaft (1985)
Entrada Sandstone Jurassic, Utah, Arizona and New Benan and Kocurek (1999)
Mexico Kocurek (1981)
Carr-Crabaugh and Kocurek (1998)
Crabaugh and Kocurek (1993)
Helsby Sandstone Formation and Triassic, NW England Herries and Cowan (1997)
Ormskirk Sandstone Formation Mountney and Thompson (2002)
Thompson (1969)
Weber Sandstone Pennsylvanian–Permian, N Utah and Driese (1985)
Colorado (Fryberger (1990d)
Stabilized eolian systems
Shikaoda Formation Proterozoic, India Chakraborty and Chakraborty (2001)
Tsondab Sandstone Tertiary, W Namibia Kocurek et al. (1999)
Tumblagooda Sandstone Silurian, Western Australia Trewin (1993)
Whitworth Formation Early Proterozoic, Queensland Simpson and Eriksson (1993)
Australia
Mixed eolian systems
Cedar Mesa Sandstone Permian, SE Utah Loope (1984, 1985)
Langford and Chan (1988, 1989)
Mountney and Jagger (2004)
Lower Cutler Beds Pennsylvanian–Permian, SE Utah Loope (1985)
Rankey (1997)
Rotliegend Group Permian, southern North Sea George and Berry (1993, 1997)
Glennie (1990)
Glennie and Buller (1983)
Howell and Mountney (1997)

Conceptual and modeling studies


Bed sets and bounding surfaces Brookfield (1977)
Fryberger (1993)
Porter (1986)
Rubin and Hunter (1982, 1983)
Rubin (1987a)
Supersurfaces, sequence Blakey (1988b)
stratigraphy and controls on eolian Kocurek (1988, 1999)
systems Kocurek and Havholm (1993)
Kocurek and Lancaster (1999)
Loope (1985)
Mainguet and Chemin (1983)
Wilson (1971, 1973)
22 NIGEL P. MOUNTNEY

sensed satellite data (Breed and Grow, 1979; Breed et al., 1979). in determining the extent to which eolian accumulations may be
The state of knowledge at the end of the 1970s enabled simple 3D preserved. Clemmensen et al. (1994) demonstrated the occur-
eolian facies models to be proposed, which, although based on rence of separate erg sequences in the ancient record, accumu-
observations from a limited number of modern and outcrop lation and preservation of which they infer to have been con-
studies, provided a basis for interpretation and a norm for com- trolled by cyclic changes in climate within gradually subsiding
parison. basins. Kocurek (1999) presented the culmination of this re-
search as a set of rules that set out the steps required for (1) erg
The 1980s: Order Out of Chaos construction, (2) the accumulation of a body of strata, and (3)
preservation of the body of strata.
Research in the early 1980s focused on the application, testing, Throughout the 1990s, researchers documented the response
and refinement of existing models through the collection of of eolian systems to Quaternary climate change. Alsharhan et al.
stratigraphic datasets from outcrop (e.g., Kocurek, 1981). Addi- (1998) and papers therein explore how eolian systems respond to
tionally, studies of modern eolian systems highlighted that the external forces such as changes in precipitation, the level of the
relatively small proportion of facies within interdune regions groundwater table, wind velocity, and, in the case of coastal ergs,
were often key to environmental interpretation (Ahlbrandt and changes in sea level. One important general conclusion from this
Fryberger, 1981; Hummel and Kocurek, 1984). Rubin and Hunter work has been the recognition that most modern eolian dune
(1982) outlined a model to explain the mechanics of eolian accu- systems are currently still responding to late Quaternary climate
mulation through the process of bedform climbing. Considerable changes and are therefore out of equilibrium with the current
research focused on an assessment of the environmental signifi- climatic regime (Lancaster, 1998). The application of geophysical
cance of eolian bounding surfaces and in particular whether techniques such as ground-penetrating radar (GPR) has enabled
major surfaces of great lateral extent were the product of regional the detailed internal architecture of modern bedforms to be
deflation to the water table as originally suggested by Stokes established in more detail than ever before and is helping to
(1968). Kocurek (1984) and Rubin and Hunter (1984) argued that establish the exact mechanisms by which modern dunes generate
surfaces arising from both climbing bedform migration and sets of strata (Bristow et al., 1996; Bristow et al., 2000). Meanwhile,
regional deflation could occur, and Talbot (1985) demonstrated new dating techniques, such as optically stimulated lumines-
regionally extensive deflationary “supersurfaces” to be a product cence (OSL) are helping to determine the migration histories of
of climatic change. Kocurek (1988) proposed that “supersurfaces” large dunes, under both present-day wind regimes and during
of regional extent originated through a variety of mechanisms previous windier glacial episodes (Bristow et al., in 2005).
that usually resulted in the termination of erg accumulation;
hence, such surfaces effectively bounded separate erg sequences MODERN AND RECENT DESERT EOLIAN SYSTEMS
in the ancient record. The development of 3D geometric strati-
graphic modeling techniques (Rubin, 1987a), which enabled the Eolian Systems
simulation of the generation of complex bed-set geometries,
provided a means to explore how relatively common styles of Approximately 30% of the present-day land surface of the
bedform migration could generate the complex bed-set architec- Earth is characterized by arid or semiarid climatic conditions
tures observed in outcrop. For example, these new techniques (Fig. 1), and eolian sand deposits cover 20% of these regions.
helped in the recognition that accumulations of linear (longitudi- Eolian processes occur preferentially in arid regions because
nal) dunes were apparently underrepresented in the rock record low amounts of precipitation result in a dry substrate with
because such bedforms usually have an additional small compo- relatively sparse vegetation cover, which promotes the ability of
nent of transverse motion, and it is this signature that is preferen- the wind to entrain and transport loose surface material. Addi-
tially preserved as sets of strata (Rubin and Hunter, 1985). By the tionally, extensive eolian processes also occur in humid settings
end of the 1980s 3D eolian facies models had been developed for where there is a surfeit of sediment supply available for trans-
specific types of eolian dune systems, including those character- port and the wind velocity is sufficiently great, the main ex-
ized by transverse, oblique, and longitudinal bedforms. ample being sandy coastlines (Pye, 1983). Most of the world’s
larger deserts can be divided into a series of distinct geomorphic
The 1990s and Beyond: elements (Fig. 2). Those elements characterized by eolian sand
Eolian Sequences and External Controls accumulation are called sand seas or ergs and range in size from
a few km2 to 560,000 km2 in the case of Rub al Khali erg, Arabia
In the 1990s, a conceptual framework was developed that (Wilson, 1973). Ergs represent the depositional part of eolian
enabled the construction of dynamic facies models that ac- systems, systems that are additionally composed of areas where
counted for spatial and/or temporal variability in controlling eolian sediment supply is generated and where eolian transport
parameters. Kocurek and Havholm (1993) discussed how sedi- and erosion occurs.
ment flux within an eolian system controls whether the sedi-
ment body undergoes accumulation, bypass, or deflation. Ad- Eolian Sediment Entrainment
ditionally, this work proposed that erg systems could be classi-
fied as dry, wet, or stabilized depending on the nature of the Eolian sediment entrainment occurs wherever and whenever
agent that controlled accumulation. Fryberger (1993) provided wind velocity is sufficient to overcome the effects of surface
a valuable review of eolian bounding surface types and summa- stabilization. At a grain scale, aerodynamic lift and drag promote
rized the terminology used in their description and interpreta- entrainment, whereas particle weight, friction, and cohesive in-
tion. Kocurek and Lancaster (1999) argued that the construction terparticle forces retard entrainment (Fig. 3). Drag and lift are
of ergs is dependent on the generation of a sediment supply, the generated by fluid flow over and around the particle, whereas
availability of that supply for eolian transport, and the ability of weight, friction, and cohesion are determined by particle size,
the wind to transport that sediment and build an erg. George density, shape, packing, moisture content, mineralogy, and de-
and Berry (1997) and Howell and Mountney (1997) demon- gree of interparticle cementation (Lancaster, 1995). Where a fluid
strated that rate of creation of accommodation plays a major role flows over a solid surface, a boundary layer develops because of
EOLIAN FACIES MODELS 23

Turkmenistan
Taklamakan
Great
Basin
Syrian Gobi
Sonoran Turpan
Mojave
Libyan Thar
Sahara Nubian
Arabian
Danakil
Ethiopian
Peruvian
(Loma)
Great
Sandy
Atacama Namib Gibson
Monte
Australian
Kalahari Simpson
Great
Victoria
Patagonia

FIG. 1.—Distribution of the world’s major climatic deserts.

friction close to the interface, such that the velocity profile in- saltation cloud within a few seconds (Fig. 5). Mid-air collisions
creases from zero at the surface itself to approach the mean wind between saltating grains allow individual grains to be held aloft
velocity some distance above the surface. For turbulent flows (as within a saltation cloud for several seconds, during which time
is typical for wind), flow mixing in the boundary layer, owing to grains may be transported tens of meters. Once saltation has been
efficient momentum transfer, results in a steep velocity gradient initiated, the energy associated with incoming grain collisions
immediately above the surface and, hence, greater shear stress means that further entrainment and transport will continue if the
(Fig. 3; Bagnold, 1941). Thus, turbulence provides a mechanism wind velocity drops below the fluid transport threshold until a
by which initial grain entrainment into the airflow can occur at lower impact threshold shear velocity is reached (Fig. 3; Bagnold,
relatively low wind velocities. 1941). Reptation is a mode of eolian transport intermediate be-
tween saltation and creep whereby larger grains hop short dis-
Eolian Sediment Transport tances downwind as they are impacted by incoming saltating
grains (Fig. 4). Creep, saltation, and reptation are all types of
Particles of larger grain size (coarse sand, granules, and rarely bedload transport. Finer-grained particles (clay and silt) are usu-
pebbles) may be transported by the wind but are usually re- ally transported as suspended load (Fig. 5), in which fine dust may
stricted to surface creep (Fig. 4), whereby grains roll or slide along be held aloft indefinitely by atmospheric turbulence and silt-
the bed during periods of high wind velocity. Creep may addi- grade material (loess) may be carried 30–300 km downwind
tionally be promoted by grains already in motion falling out of the during single wind-storm events (Pye and Tsoar, 1987).
airflow and nudging surface grains downwind. Grain size, shape,
density, sorting, and packing all determine the ease with which Eolian Sediment Textures
grains undergo creep. Particles of very fine to medium sand,
which constitute the bulk of sediment transported in most eolian Eolian sediments often exhibit distinctive grain size, shape,
systems, are susceptible to saltation, whereby grains are lifted into and sorting characteristics that can be useful in the recognition of
the airflow and carried downwind before returning under the eolian strata (Pye, 1982; Lancaster, 1986). The wind is highly
influence of gravity to the surface (Fig. 4). Incoming grains often selective in terms of the grain sizes that it can carry for a given
possess sufficient energy to bounce (saltate) back into the airflow. velocity, and eolian dunes are often characterized almost exclu-
Importantly, saltating grains additionally impact and dislodge sively by very fine- to medium-grained sand that is well or very
other grains on the bed (ejecta), causing them to commence well sorted (Fig. 6). Interparticle collisions result in high rates of
transport and thus inducing a chain reaction that generates a grain abrasion, and less resistant minerals such as mica and lithic
24 NIGEL P. MOUNTNEY

N
50 km

Turkmenistan
Taklamakan

Gobi
Turpan
Sahara Thar

Arabian

FIG. 2.—Satellite image of the Turpan Depression, Bogda, northern China. This arid system contains a number of erosional and
depositional elements including eroding mountain catchment areas, alluvial fans, wadi channels, an interior-draining salt lake,
salt flats (inland sabkha), sand sheets, and a major erg accumulation. Image courtesy of NASA Earth Observatory collection.
EOLIAN FACIES MODELS 25

A L
L = lift
D = aerodynamic drag
Wind D W = weight
C = inter-particle cohesion.

C C

B 40 40
Turbulent
30 30 flow
Height (mm)

Height (mm)
20 20

10 Laminar 10
flow
0 0
0 1 2 3 0 1 2 3
Wind velocity (m/s) Wind velocity (m/s)

C 80
Higher wind speeds Fluid
needed for small threshold
Wind velocity (m/s)

60 grains due to
cohesive forces

40
Impact
threshold
20

0
0 0.04 0.08 0.4 0.8 1.2 1.6
Grain diameter (mm)

FIG. 3.—Factors governing particle entrainment by the wind. A) Forces exerted by the wind on a particle at rest. B) Typical vertical
velocity profiles showing the smaller near-surface shear stresses in laminar flow when compared with turbulent flow. C)
Relationship between particle size and threshold shear velocity. Saltating grains lower the wind speed needed to induce further
grain motion. After Bagnold (1941).

fragments often break down to dust, whilst more resistant grains position and texture is ultimately dependent on the nature of the
(e.g., quartz) become highly rounded as angular edges are abraded, source material, the availability of that material for eolian trans-
often develop a high sphericity, have surfaces that are dull port, transport distance, and wind gustiness. Carbonate eolian
(frosted) as a result of repeated grain collisions, and may exhibit dunes are documented from Oman and the United Arab Emirates
conchoidal fractures. Eolian grains with these properties often (Besler, 1982), and eolian dunes composed of basalt and pumice
have a “millet seed”’ texture (Fig. 6). Although many eolian sands clasts have been recorded from active volcanic provinces
are monomineralic (usually dominated by quartz grains), com- (Mountney and Russell, 2004). The movement of coarse sand and
26 NIGEL P. MOUNTNEY

A 0.5–2.0 mm

Lift Impact Wind Impact

Drag

C Saltating grain
Reptating grain

D W1 Final
Initial U2
W2
U1 Saltation
b height
a

Saltation path length

FIG. 4.—Methods of eolian grain transport. A) Surface creep. B) In-air collisions of saltating grains maintains momentum, keeping
grains aloft. Ground impacts induce new grains to saltate. C) Impact of saltating grains with grains on bed drives reptation. D)
The ballistic trajectory of a saltating sand grain. W and U represent vertical and horizontal velocities, respectively. a is the approach
angle, b is the take-off angle.

granules via creep typically occurs only sporadically during serir). The concentration of larger clasts on deflationary surfaces
episodes of high winds, and sand-sheet deposits where creep is acts as an armored lag that protects underlying sand from further
a dominant process are often characterized by poorly to moder- winnowing (Fig. 6D). Sand sheets often contain wind-faceted
ately sorted sediments. cobbles and boulders (ventifacts) with distinctive upwind-facing
Eolian deflation involves the winnowing of sand-grade sedi- scalloped and abraded surfaces that are useful paleocurrent
ment to leave a surface dominated by a coarse-grained lag (reg or indicators.
EOLIAN FACIES MODELS 27

and Iverson, 1985; Fryberger et al., 1992) and wavelengths and


A heights up to 5 m and 0.35 m, respectively (Fig. 8D). A continuum
of ripple sizes exists between these limits (Ellwood et al., 1975).
Eolian ripples can be differentiated from subaqueous ripples
because the former typically have high ripple indices (ratio of
wavelength to height) of 25–40+, and are often characterized by
inverse grading that results from the migration of coarser-grained
ripple crests over finer-grained ripple troughs (Fig. 9). The major-
ity of eolian ripples develop as a consequence of saltation and
reptation. For a given wind velocity, grains in motion are re-
stricted to a narrow size range, and the distance that grains jump
downwind (saltation path length) is similar for most of the sedi-
ment in transport (Bagnold, 1941). Eolian saltation ripples begin
to form with spacings that are determined by the saltation path
length (Figs. 4D, 9). Minor surface perturbations act as the catalyst
required to initiate ripple development (Anderson, 1987), and,
B once initiated, the ripples themselves grow and steepen into
bedforms because upwind-facing stoss slopes act as an impact
zone that catches incoming saltating grains, whereas downwind-
facing lee slopes act as a shadow zone where grain impacts are
minimal (Sharp, 1963). Grains landing in the impact zone often
creep up the stoss slope to the ripple crest before once again being
launched into the airflow and saltating downwind to the next
ripple. Coarser grains often concentrate at ripple crests, whereas
finer grains are preferentially trapped in ripple-trough shadow
zones.

Dunes.—
Eolian dunes have wavelengths of 5–250 m (Fig. 7) and are
often arranged into trains of regularly spaced bedforms (Lancaster,
1988; Werner and Kocurek, 1999). Most dunes have a windward
stoss slope inclined at 8–16° and a lee slope inclined at 20–34°. Dunes
FIG. 5.—Examples of eolian grain transport. A) Saltation of sand- form topographic obstacles that disrupt the primary airflow such
size particles across a low-relief sand sheet. Most transport that as the flow moves up the dune stoss slope it accelerates,
occurs within 2 m of the surface. Skeleton Coast, northern thereby causing an increase in transport rate and promoting
Namibia. B) Suspension of silt-size particles within the air- transport up the stoss slope to the dune crest. As the flow moves
flow to a height in excess of 300 m above the surface. Huab over the crest and into the lee-side depression, it decelerates, and
Basin, northern Namibia. causes a decrease in transport rate, thus promoting deposition on
the lee slope (Sweet and Kocurek, 1990; Frank and Kocurek,
1996). This provides the basis for a mechanism by which dunes
Although textures can aid the recognition of ancient eolian advance downwind over time. Flow separation of the airflow from
deposits, they can be reliably used only when associated with the bedform surface occurs beyond the crest, whilst flow reattach-
diagnostic sedimentary structures. The intimate association of ment typically occurs a distance of about seven dune heights
many eolian systems with fluvial, lacustrine, and coastal environ- downwind. Thus, a separation cell exists in the dune lee (Sweet,
ments means that sediments with eolian textures are frequently 1992) within which turbulent secondary airflow occurs that allows
reworked by non-eolian processes. For example, the Kuiseb River ripples and erosional scour hollows on the dune flanks (plinth) to
of Namibia captures eolian sand moving north from the Namib undergo complex migratory behavior. Downwind of the reat-
Sand Sea and transports it downstream before releasing it to the tachment point, renewed flow acceleration means that interdune
shallow offshore realm. sediments may potentially be eroded, thereby providing a local
sediment supply for the next dune downwind in the train. The
Eolian Bedforms angle of repose for most types of dry eolian sand is 32–34°, and lee
slopes inclined at or beyond these angles are inherently unstable
Eolian bedforms can be classified according to their scale, and develop an active slipface where grainflow avalanche pro-
morphology, orientation relative to net sand transport direction, cesses dominate (Hunter, 1977), whereas lee slopes inclined at
style of migratory behavior, and style of superimpositioning. less than 32° are generally stable and characterized by ripples.
Three distinct scales of eolian bedform are recognized: ripples, Dunes exhibit a wide variety of morphological forms (Figs. 10,
dunes, and draa (Fig. 7). 11) that reflect the combined effects of a number of controlling
factors, including wind strength and directional variability on
Ripples.— diurnal to seasonal (and longer) timescales, the timing of genera-
tion of a sediment supply, and the availability of that sediment for
Eolian sand ripples typically have straight crestlines oriented transport. Dunes can be classified as mobile (actively migrating),
perpendicular to wind direction, and have wavelengths of 50–200 active but anchored, or stabilized (Fig. 12; Cooke et al., 1993).
mm and heights of 5–10 mm (Fig. 8; Bagnold, 1941). Coarser Mobile dunes can be classified according to their morphology on
granule megaripples typically have sinuous crestlines (Greeley the basis of the number of lee faces that they possess (McKee,
28 NIGEL P. MOUNTNEY

A B

C D

FIG. 6.—Eolian sediment textures. A) Well rounded and well sorted “millet seed” grains. B) Eolian sand composed of frosted quartz
grains and lithic fragments. C) Bimodally sorted sand on an eolian ripple (coarse grains on crest). D) Pebble deflation lag.

1979), and according to the orientation of their crestlines relative particular dune types (Fig. 13). Drift potential (DP) is a measure
to the predominant wind direction (Hunter et al. 1983). Common of the total-sand moving capability of the wind without regard
dune types classified according to these criteria include transverse to wind direction. Resultant drift potential (RDP) is a measure of
dunes, which have a single lee face and a crestline normal to the the resultant or net sand-moving capability of the wind in the
prevailing wind, linear or seif dunes, which have one or two lee resultant drift direction (RDD). RDP/DP, the unidirectionality
faces and a crestline parallel to the wind, and star (pyramid) dunes, index of Wilson (1971), is a measure of wind variability where
which have three or more lee faces (Fig. 10). The classification of values approaching unity (RDP/DP > 0.8) signify low variabil-
dunes as transverse or longitudinal is potentially misleading ity and low values (RDP/DP < 0.3) signify high variability. One
because net sand transport direction across many dunes is oblique of the original methods for the classification of dune types was
to the orientation of the bedform (Hunter et al. 1983; Rubin and based on the number and orientation of slipfaces (McKee, 1979),
Hunter, 1985). At a more detailed level, dunes that possess straight which in turn reflects the complexity and variability of the wind
crestlines are two-dimensional, whilst those with sinuous, cuspate, responsible for generating and maintaining the bedform. Trans-
or lobate crestlines are three-dimensional (Rubin 1987a). Isolated verse dune forms tend to develop under conditions of unidirec-
barchan dunes and barchanoid dune ridges are examples of 3D, tional winds characterized by high RDP/DP values and are
transverse bedforms. The style of migratory behavior of mobile sand-transporting bedforms, whilst star-dune forms tend to
dunes can also be used for classification; dunes that migrate in a develop in response to variable winds (low RDP/DP values)
constant direction, at constant speed, and without undergoing (Wasson and Hyde, 1983) and are sand-accumulating bedforms
charges in form over time are invariable, whereas dunes that that do not migrate great distances (Fig. 13).
undergo temporal changes in migration direction, speed, asym-
metry, and/or steepness are variable (Rubin, 1987a, 1987b). Draa.—
Fryberger (1978, 1979) and Fryberger et al. (1979) defined
three terms that attempt to classify the energy and directional Draa are larger-scale bedforms than dunes (Wilson, 1971,
properties of the wind and relate it to the construction of 1972) that have wavelengths of 500–5000 m and exceed 50 m in
EOLIAN FACIES MODELS 29

20 RIPPLES
Grainsize (mm)
10
DUNES
8
6 DRAA
4

1 4 16 64 256 10 40 160 640 2560

centimeters meters

Bedform wavelength

FIG. 7.—Grain size (coarsest twentieth percentile) versus wavelength for eolian bedforms. Note the three distinct groups representing
ripples, dunes, and draa. Modified after Wilson (1972).

A B

C D

FIG. 8.—Examples of eolian ripple forms. A) Sinuous crested with coarser grains on crests. Skeleton Coast, Namibia. B) Two scales
of superimposed ripples. Idaho (courtesy of John Collinson). C) Two scales of ripples developed on the stoss slope of an eolian
dune. Huab Basin, Namibia. D) Sinuous-crested eolian granule megaripples. Askja sandsheet, central Iceland.
30 NIGEL P. MOUNTNEY

A B C
Variation in impact intensity over pertubation in bed.
Note higher intensity in AB compared to BC

Ripple spacing is controlled by saltation path length, which


is itself primarily a function of grain size and wind velocity.

Impact
angle

Impact Shadow Impact


zone zone zone
Alternation of impact and shadow zones on developing
wind ripple (after Anderson, 1987).

B Wind direction
Crestal accumulation
of coarser grains
Veneer of
finer grains

Core of finer grains Foreset bed


Laminae and grain size distribution within eolian ripples (after Sharp, 1963).

FIG. 9.—Generation of eolian ripples. A) Model for the generation of saltation ripples. After Anderson (1987). B) Grain texture in
eolian ripples. After Sharp (1963).

height (Fig. 7). These “mega-bedforms” occur only in the largest recognized to occur in various configurations and often combine
ergs, where eolian sediment supply and transport rates are high. to form larger-scale cross-bedded sets.
Draa are described using the same terminology as for dunes but
additionally can be characterized by the presence of superim- Wind-Ripple Strata.—
posed dune-scale bedforms on their flanks (Fig. 14). Simple draa
lack superimposed bedforms, whilst compound draa possess su- Tractional processes that generate wind ripples give rise to
perimposed dunes of the same morphological type and complex various types of wind-ripple stratification (Fig. 15). Although ripple
draa possess superimposed dunes of a different type (McKee, foreset laminae (rippleform laminae) may be preserved in wind-
1979). The migration of superimposed dunes over larger, more ripple strata, internal laminae often cannot be distinguished
slowly moving draa is one possible explanation for the origin of because of the uniformity of the grain size, and translatent rippleform
geometrically complex bed sets in the ancient record. stratification results (Fig. 16; Hunter, 1977). Wind-ripple strata
sometimes exhibit a weak inverse grading, in part because the
Eolian Sedimentary Structures and Stratification finest material tends to accumulate in sheltered ripple troughs,
whilst the upper parts of ripples are composed of coarser material
Small-scale eolian stratification arises in response to a distinct (Fig. 17), and in part because finer grains tend to settle between
suite of processes that enables eolian strata to be recognized coarser grains, resulting in a pour-in texture (Sharp, 1963; Fryberger
(Hunter, 1977; Hunter, 1981; Kocurek and Dott, 1981; Fryberger et al., 1992). This means that the base of the ripple stratum is often
and Schenk, 1981). Four basic eolian stratification types (ripple distinct, the surface being defined by a thin lag of finer material.
strata, avalanche strata, grainfall strata, and adhesion strata) are Where ripples preserve ripple traces only one or two grains thick,
EOLIAN FACIES MODELS 31

A barchan B barchanoid

C transverse D oblique

E longitudinal F parabolic
(linear, seif)

G star H dome
(pyramid)

FIG. 10.—Three-dimensional forms of some common dune types. The arrows mark the dominant directions of the effective winds and,
in case E, the dotted arrow indicates the resultant effective direction.

a characteristic pinstripe lamination is preserved (Fig. 15A; Fryberger generates a scarp that then retreats back upslope toward the
and Schenk, 1988). Eolian ripple strata form widespread deposits brinkline. The resultant deposits form tongue-like bodies that
in sand sheets, on dry interdunes, and on low-moderately in- rarely exceed 0.5 m in width but may extend almost the full length
clined dune and draa slopes. of the lee slope. Grainflows may exhibit inverse vertical grading
due to shear sorting and downslope grading where the coarser
Grainflow Strata.— sediment grains run farthest down slope. Where developed in
very well sorted sand, the boundaries between successive ava-
When the lee slope of an eolian dune exceeds the angle of lanches might not be evident, in which case only amalgamated
repose (32–34°), an active slipface develops that is subject to grainflow units will be recognized (Howell and Mountney, 2001).
gravity-driven collapse, resulting in the generation of various
types of avalanche strata (fig. 18). Lee-slope avalanches are ex- Grainfall Strata.—
amples of grainflows (sandflows) in which intergranular cohesion
is lost during the flow, resulting in an erosionally based, chaotic Gravity-driven grainfall occurs as the wind carries clouds of
deposit that usually lacks internal structure. Slump degradation saltating grains over a dune brink. A reduction in wind transport
grainflows occur where internal structure is destroyed as the flow capacity in the lee-side depression allows grains to settle onto the
travels downslope, and resultant deposits are characterized by a upper part of the lee slope (Fig. 19; Nickling et al., 2002). Grainfall
chaotic wedge of loosely packed sediment up to a few meters strata are usually difficult to distinguish but are often moderately
wide that thickens downslope up to 5–6 cm before pinching out. packed and exhibit a wedge-shaped geometry that is thickest just
Scarp recession grainflows occur where an initial point of failure leeward of the brinkline and thins downslope (Hunter, 1985;
32 NIGEL P. MOUNTNEY

A B

C D

FIG. 11.—Examples of eolian dune forms. A) Slipface and plinth of crescentic barchan dune. Skeleton Coast, Namibia. B) Transverse
dunes. Western Namib Sand Sea. C) Linear dune ridge partly stabilized by vegetation. Lake Eyre Basin, Australia (courtesy of
John Collinson. D) Large star dune, central Namib Sand Sea.

Anderson, 1988). Additionally, grainfall strata often blanket the Cross Bedding.—
upper parts of dune lee slopes for distances of tens of meters
along-slope (Hunter, 1977), enabling them to be differentiated Cross bedding is ubiquitous within eolian dune sands and
from individual avalanche stratum. On small dunes, wedges of sandstones. It develops as a consequence of repeated and ongo-
grainfall strata may extend down to the base of the lee slope, ing lee-slope sedimentation whereby ripple, avalanche, and
whereas on dunes of increasing size, wedges of grainfall strata grainfall strata generate cross stratification (Fig. 21). The interiors
tend to be cut out by grainflow strata. Repeated grainfall deposi- of most eolian bedforms are composed of cross-bedded sands,
tion on the upper lee slope is the main mechanism by which the and the stratification planes provide a record of the former
slope attains and exceeds the angle of repose, thus inducing positions and shape of the bedform lee slope and of the processes
reworking of grainfall strata by avalanche processes. that operated on that slope (Figs. 22, 23). Where bedforms migrate
over one another, cross strata are truncated and sets delineated by
Adhesion Strata.— erosional bounding surfaces are generated.

The adhesion of grains in motion to a damp surface results in Structures Due to Postdepositional
the generation of a range of structures (Fig. 20) including adhe- Soft-Sediment Deformation.—
sion plane beds, adhesion ripples (Kocurek and Fielder, 1982),
and adhesion warts (Olsen et al., 1989), which are characterized Small-scale deformation structures (< 1 m) in eolian strata
by low-relief ridges and mounds that grow by adhesion to their occur as either intradune folding, indicative of near-surface liq-
upwind edge and thereby undergo upwind migration. Adhesion uefaction, or slumping of moderately cohesive, moist sands on
structures are preserved both on bedding surfaces and in section the dune lee slope in response to surface precipitation (Doe and
where strata form crinkly and wavy laminae. The generation of Dott, 1980). Liquefaction results from an elevation of pore-water
adhesion strata requires the accumulation surface to be damp, pressure as the wetting front infiltrates into highly porous dune
and such strata often occur in low-lying damp interdune and sands (Fig. 24). Loosely packed grainflow laminae are particu-
dune-flank settings (Hummel and Kocurek, 1984). larly susceptible to liquefaction by collapse of grain packing due
EOLIAN FACIES MODELS 33

A Mobile B Anchored

Type Description Type Description

TRANSVERSE Asymmetrical, crestline normal RISING Large sand banks that


to transport direction, single form on windward side of
slipface, unidirectional wind topographic obstacles.

Barchan Isolated crescentic dune FALLING Large sand banks that


form on leeward side of
Barchanoid Sinuous-crested ridge topographic obstacles.

Transverse Straight-crested ridge ECHO Dunes on steep windward slope


of topographic ridge but separated
LINEAR (SEIF) Symmetrical, crestline parallel from ridge by sand-free area due
to transport direction, often with to occurrence of a fixed eddy.
2 slipfaces, bidirectional wind
SHADOW Small sand dunes that form
Straight Straight-crested ridge in localized topographic
depressions or in lee of obstacles
Sinuous Sinuous-crested ridge,
scour pits migrate along flanks FLANK Dunes anchored to the flanks
of larger parent forms. Result
STAR Peaked form with 3+ slipfaces from secondary airflow.
separated by ridges or spurs.
Multidirectional wind regime.

Pyramid Central peak, no elongate arms C Stabilized


Radiating Central peak, with elongate arms
Type Description
SAND SHEET Low-relief sand accumulation
that actively accumulates but NEBKHA Sand accumulation developed
lacks slipfaced bedforms. around vegetation clumps. Also
known as coppice dunes.
Zibar Low-relief dunes that lack
slipfaces, varied morphology PARABOLIC U-shape dunes with active “nose”
and trailing arms stabilized by
SUPERIMPOSED Transverse, linear and star vegetation. Various scales up to
parent forms that support smaller 1–2km long and 10–70m high.
dunes on their flanks. Often
large scale (draa). CEMENTED Dunes that undergo early cement-
ation due to the precipitation of early
Compound Superimposed dunes of same diagenetic cement (e.g. gypsum),
morphological type as parent often associated with moisture.

Complex Superimposed dunes of different BEACH DUNE Coastal backshore dunes stabilized
morphological type to parent RIDGE by vegetation. Often developed in
humid climatic settings

FIG. 12.—Classification scheme for common dune types. A) Mobile (freely migrating) dunes, B) anchored dunes that are active but
fixed because of aerodynamic conditions, C) stabilized dunes that develop because of the action of physical, chemical, or biogenic
factors favorable for sand accumulation. The scheme is not rigid, and overlap exists between the three groups. For example,
nebkha and parabolic dunes could be classed as being anchored by vegetation. Many dunes are neither transverse nor linear but
are oriented oblique to resultant sand transport direction.

to mechanical loading associated with the wetting event. Defor- Fryberger, 1981; Kocurek, 1981). Sedimentary structures of physi-
mation due to surface collapse occurs where the upper dune lee cal origin (Fig. 27) include desiccation cracks and polygons,
slope fails because of a decrease in the angle of internal friction in raindrop imprints, mud flakes and curls, wave and current
response to wetting and to oversteepening by wet grainfall ripples, cross strata, and wavy lamination of subaqueous origin
(Hunter et al., 1983; Loope et al., 2001), resulting in a range of (Ahlbrandt et al., 1978; Langford 1989; Langford and Chan,
brittle-failure structures, including slab slides (Fig. 25). Larger- 1989). Sedimentary structures of chemical origin (Fig. 28) in-
scale deformation structures usually indicate liquefaction below clude evaporitic salt crusts, fenestral porosity, sandstone pseudo-
the water table (McKee et al., 1971), which is usually best ex- morphs of salt minerals, especially halite (hoppers) and gypsum
plained by loading of the saturated sand by an advancing dune (desert rose), and evaporite precipitation structures such as tee-
(Fig. 26; Collinson, 1994; Horowitz, 1982). pees (Kocurek, 1981). Sedimentary structures of biogenic origin
(Fig. 29) include animal footprints, trackways, burrows and
Non-Eolian Sedimentary Structures.— crawling traces (Ahlbrandt et al., 1978; Hasiotis, 2002), plant
root structures and rhizoliths (Loope, 1988), and algal growth
Non-eolian processes and structures are common in many structures. Paleosols are also widespread in many desert sys-
eolian systems, especially in interdunes (Ahlbrandt and tems (Kocurek et al., 1991a).
34 NIGEL P. MOUNTNEY

A DP = 518 DP = 255
RDP = 448 RDP = 56
RDD = NE RDD = SSE
RDP/DP = 0.86 RDP/DP = 0.22

Crescentic dunes Complex star draa


Walvis Bay, Namibia Ouargia, Algeria
Narrow unimodal Complex distribution

B 50
sand thickness in meters)
Amount of sand in dunes
(expressed as average

40
Star
Transverse
30

20
Linear
10 (Longitudinal)
Crescentic
(Barchan)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Wind-directional variability (RDP/DP)

FIG. 13.—Illustration of the concept of drift potential (DP), resultant drift potential (RDP), and resultant drift direction (RDD). A) Sand
transport regimes represented by sandflow roses (after Fryberger, 1979). All numeric values are in vector units. See text for further
explanation. B) Relationship between dune type, wind regime, and equivalent sand thickness. Transverse and barchan dunes
develop under unimodal wind regimes (RDP/DP > 0.5) and are sand-transporting bedforms. Star dunes develop under multi-
directional wind regimes (RDP/DP < 0.2) and are sand-storing bedforms. Modified after Wasson and Hyde (1983).

Eolian Bounding Surfaces and Cross-Bedded Sets Reactivation Surfaces.—


Bounding surfaces are erosional surfaces that are generated Reactivation surfaces result from periodic lee-slope erosion fol-
as an intrinsic product of eolian dune migration whereby lowed by renewed sedimentation associated with a change in
bedforms (or parts thereof) scour into preexisting deposits as bedform migration direction, migration speed, asymmetry, and/
they move through space (Fig. 30). Brookfield (1977) produced or steepness (Fig. 30; Rubin, 1987a, 1987b). These changes are
the first widely accepted model that explained a hierarchy of common because airflow on lee slopes is often subject to turbulent
eolian-bounding surface types (first-, second-, third-order) in modification and is rarely steady. In some cases, the period of the
terms of the migratory behavior of bedforms (Fig. 31). Subse- flow fluctuation is regular and generates cyclic reactivation sur-
quently, Rubin and Hunter (1983) and Hunter and Rubin (1983) faces, as is the case for diurnal and seasonal wind reversals (Hunter
argued that apparently similar geometrical bounding-surface and Rubin, 1983; Hunter et al., 1983; Hunter and Richmond, 1988;
configurations could be produced by a variety of styles of Loope et al., 2001). Nested reactivation surfaces on two or more
bedform behavior, and Rubin (1987a) employed 3D computer scales occur when cyclic cross bedding is generated by the interac-
simulations to demonstrate how arrangements of bedforms tion of two or more forcing parameters operating with different
could generate highly complex bounding-surface geometries. periodicities (Fig. 32; Crabaugh and Kocurek, 1993).
Fryberger (1993) provides a valuable summary of terminology Reactivation surfaces occur within eolian sets and are char-
for eolian bounding surfaces and provides some useful photo acterized by planar- or scalloped-shaped erosion surfaces that
examples. Three broad types of eolian bounding surface are typically dip downwind at inclinations of 10–20°, somewhat
recognized to occur as a product of autocyclic (intrinsic) bed- less than the cross strata that they truncate (Fig. 30). In sections
form migratory behavior. perpendicular to eolian transport, reactivation surfaces trend
EOLIAN FACIES MODELS 35

B C

FIG. 14.—Examples of eolian draa forms. A) Pyramid-shaped star draa with radiating linear ridges, some supporting superimposed
transverse ridges. Central Namib Sand Sea. B, C) Complex linear draa with numerous superimposed bedforms, mostly oriented
oblique to the trend of the main crest line. Northern Namib Sand Sea.

parallel to subparallel to the cross strata and can sometimes be Superimposition Surfaces.—
traced for 10–100+ m along strike, whilst in sections parallel to
transport they may extend the full height of a set or may be Superimposition surfaces result from either the migration of
restricted to its basal part, in which case they are often charac- superimposed dunes over a larger parent bedform, or the mi-
terized by a sweeping (asymptotic) base. Reactivation surfaces gration of scour troughs on the lee slope of a bedform (Fig. 30;
either occur randomly within sets or exhibit regular spacings, as Rubin, 1987a). Although theoretically superimposed dunes and
has been documented from the Navajo Sandstone (Chan and scour troughs can migrate directly up or down the lee slope of
Archer, 1999, Loope et al., 2000). Overlying cross strata exhibit a parent bedform, oblique migration is more common because
either a concordant or a downlapping relationship. secondary airflow, which occurs because the high-relief bedforms
36 NIGEL P. MOUNTNEY

A B

C D

FIG. 15.—Examples of eolian ripple internal stratification. A) Pinstripe lamination, Etjo Formation, Cretaceous, Namibia. B) Inversely
graded translatent strata, Askja, Iceland. C) Sharply defined wind ripple laminae interbedded with thin grainfall laminae. Lower
Cutler Beds, Pennsylvanian–Permian, Utah, U.S.A. D) Wind-ripple strata on a dune plinth. Cedar Mesa Sandstone, Permian,
Utah, U.S.A.

Translatent strata Rippleform laminae


Subcritical

Incomplete rippleform laminae


(ripple foreset cross-laminae)
(a < b)
Angle of ripple climb (a)

of ripple stoss slope (b)


relative to inclination

Subcritically climbing Truncated ripple-foreset


translatent strata cross laminae
Critical
(a = b)

Critically climbing Complete ripple-foreset


translatent strata cross laminae
Supercritical
(a > b)

Supercritically climbing Complete rippleform


translatent strata laminae

FIG. 16.—Classification of wind-ripple stratification types according to angle of ripple climb relative to the inclination of the stoss
slope of the bedform and the presence or absence of cross-lamination. Modified after Hunter (1977).
EOLIAN FACIES MODELS 37

FIG. 17.—Wind ripples generated by ballis-


tic impact of grains. The ripple spacing
A relates in a general way to the saltation
path length, which is the characteristic
distance that individual grains hop
downwind as a result of grain collision
on the bed. The saltation path length is a
function of grain size, shape and den-
sity, and mean wind velocity and gusti-
ness close to the bed. A) The migration of
wind ripples generates subparallel lami-
nation. B) The impact angle of saltating
sand grains differs between stoss sides
and lee slopes. High-angle impacts on
B the stoss of the bedforms promotes creep
high-angle of coarser grains towards the ripple crest.
impacts Downwind-facing lee slopes form a
few low-angle impacts shadow zone where relatively few low-
in shadow zone angle impacts occur, thus encouraging
the accumulation of finer grains in ripple
troughs. As ripples migrate downwind,
this sorting mechanism generates lami-
nation with inverse grading.

A B C

1m 1m

D E

FIG. 18.—Examples of eolian grainflows and the characteristic strata that they produce. A) Scarp-recession grainflow, Namibia. B)
Slump-degradation grainflows, Namibia. C) Slab slide failure degenerates downslope into a slump-degradation grainflow,
Namibia. D) and E) Grainflow tongues pinching out into wind-ripple strata. Cedar Mesa Sandstone, Permian , Utah, U.S.A.
38 NIGEL P. MOUNTNEY

A B

FIG. 19.—Examples of eolian grainfall and the characteristic strata that it produces. A) Saltation of sand-size particles over the brink
of a dune to form a suspension cloud. Deceleration of the airflow in the lee-side depression results in a loss of carrying capacity,
and the grains fall onto the upper part of the lee slope as grainfall deposits. Kalahari Desert. B) Grainfall facies interbedded with
wind-ripple strata. Individual grainfall units rarely exceed 5 mm in thickness but tend to be laterally continuous along the strike
of the cross-bedding for several meters to tens of meters. Interbedded units of wind-ripple strata are thicker (1–2 cm). Cedar Mesa
Sandstone, Permian, Utah, U.S.A. Penknife for scale.

disrupt the airflow, tends to be directed along the lee slope of the that immediately overlie an interdune surface are determined by
parent bedform. the processes operating within the interdune. For example, re-
Superimposition surfaces occur within eolian co-sets (Fig. 30) stricted interdune hollows are typically characterized by dry wind-
and are characterized by planar to highly scallop-shaped erosion ripple strata, damp interdune flats by adhesion strata, and wet
surfaces that dip in a wide range of orientations. In sections interdune ponds by subaqueous (non-eolian) strata.
parallel to transport, these surfaces appear similar to reactivation Interdune migration surfaces are characterized by low-angle-
surfaces, and their identification can be problematic. However, in inclined erosional surfaces (Fig. 30) that typically extend down-
sections perpendicular to transport, superimposition surfaces wind for distances of hundreds of meters to several kilometers.
differ from reactivation surfaces because they are usually ori- These surfaces, which bound sets or co-sets, appear planar to
ented oblique to the cross strata that they truncate. Where both slightly scalloped-shaped in sections parallel to eolian transport,
reactivation and superimposition surfaces are developed, the whilst in sections perpendicular to transport they may be moder-
latter always truncate the former. ately to highly scalloped (Kocurek, 1981; Mountney and Howell,
2000). Interdune surfaces truncate both superimposition and
Interdune Migration Surfaces.— reactivation surfaces.

Interdune migration surfaces result from the migration of bedforms Relation between Bed Sets,
separated by interdunes (Fig. 30). The surfaces are carved by the Bounding Surfaces, and Bedforms.—
passage of an erosional scour that defines the interdune trough
between successive bedforms. The depth to which the interdune Relating the morphology and migratory behavior of modern
trough scours as it migrates influences the extent to which deposits bedforms to the architecturally complex bed-set and bounding-
of the preceding bedform are eroded. The nature of the deposits surface geometries that they generate is an important compo-
EOLIAN FACIES MODELS 39

A B

C D

FIG. 20.—Examples of eolian adhesion ripples. A) Sólheimasandur, southern Iceland. Accretion occurs on the steeper upwind-facing
slopes, and the “ripples” migrate upwind over time. B) Askja, central Iceland. C) Adhesion warts, Mojave desert. D) Adhesion
structures on a bedding surface, Precambrian, Greenland (courtesy of John Collinson).

nent in the development of eolian facies models, but is far from must be performed with caution. Rubin and Hunter (1983) and
straightforward. Whilst the external morphology of modern Rubin (1987a, 1987b) document numerous examples where
eolian bedforms is readily apparent, their internal bed-set archi- mean foreset azimuths differ markedly from the migration
tecture is difficult to determine. From the geometrical computer direction of parent and/or superimposed bedforms. Such dif-
simulations of Rubin (1987a) it is clear that bedforms of similar ferences arise most commonly either from the oblique or along-
external morphologies can generate radically different patterns crest migration of superimposed dunes over parent bedforms,
of cross bedding because they undertake different migratory the oblique or along-crest migration of scour pits (Fig. 33), or the
behavior through time. Furthermore, the amount of a bedform transverse migration of scour pits that are offset relative to each
that is accumulated as a bed-set (i.e., not eroded) following the other in successive bedforms within a train such that one side of
passage of subsequent bedforms in a train is typically only a the troughs filled with cross strata are consistently eroded by
small fraction (usually < 10%) from the basal-most part of the the succeeding scour, thus preserving a “biased” record. Even
entire bedform. As such, the reconstruction of bedform mor- relatively simple bedform configurations can generate bed-set
phologies from bed-set architecture usually relies on the as- and bounding-surface architectures that are difficult to inter-
sumption that the preserved bottom sets adequately reflect the pret, and the belief that mean foreset azimuth direction is an
depositional processes that occurred on the upper (non-pre- indicator of paleodune migration direction is incorrect for all
served) parts of the bedform lee slope. but the simplest of 2D (i.e., straight crestline), invariable, trans-
verse bedforms. Furthermore, the limitations of 2D outcrop and
Reconstructing Bedform Migration Vectors from 1D core can also lead to misinterpretation because troughs need
Cross Bedding and Bounding Surfaces.— to be observed in three dimensions in order to fully appreciate
their structure (Fig. 34; DeCelles et al., 1983).
Simple facies models for “transverse” or “linear” eolian The correct method for determining the migration direction of
dune systems are misleading, and the reconstruction of parent and superimposed bedforms from foreset and bounding-
paleodune migration directions from foreset dip-azimuth data surface azimuths is discussed by Rubin and Hunter (1983) and well
40 NIGEL P. MOUNTNEY

Modern examples Eolian facies distribution on crescentic dunes Ancient examples


1 plan view 4
Stoss
slope
Grainflow
strata

Grainfall
Lee slope
slipface th
3 5 strata
plin
ne
Du
Damp Wind ripple
Slipface on barchan dune with grainflow avalanches. Wind
1 interdune strata Wind ripple strata in section displaying characteristic pinstripe
ripples in dry interdune. Skeleton Coast, Namibia. lamination. Etjo Formation, NE Namibia.
2 Adhesion
2 strata 5
Large Dune
Contorted Rotated
bedding block
Grainfall
Breccia
Grainflow
Wind ripple
6 deposits
Adhesion structures on a damp interdune surface. Plan view.
Monument Valley, northern Arizona.
4 A Wavy laminae in damp interdune unit passing up into overlying
wind ripple dune plinth strata. Helsby Sandstone, UK.

B
3 Only wind-ripple-dominated 6
basal part of underlying dune set
Small Dune preserved
Grainfall
Grainflow Shallower
truncation
A
B
Deeper
Slipface collapse due to cohesive slab slide. Slabs of wet sand Wind ripple truncation Grainflow tongues merging with wind ripple strata that
fail without loss of internal structure. Askja region, NE Iceland. deposits represent dune plinth deposits. Cedar Mesa Sandstone, Utah

FIG. 21.—Examples of characteristic eolian facies and their distribution on a simple crescentic (barchan) dune and on large-scale and
small-scale eolian dunes truncated to different levels (A, B). Level of truncation influences the preservation of facies types in the
geological record, with features characteristic of the upper slipface lost. Modified after Kocurek and Dott (1981).

illustrated by Kocurek (1996). The trend of the crestline of the and shapes, moving at varying rates and in varying directions
parent bedform is given by the strike of the interdune bounding relative to one another. In situations where the net sediment
surface, the migration direction of the parent bedform being nor- budget is positive, accumulation occurs such that bedforms climb
mal to this trend. The trend of the crestlines of superimposed over one another at various angles, thereby preserving sets of cross
bedforms is determined most easily using a stereonet. The line of strata and associated bounding surfaces (Brookfield, 1977; Rubin
intersection between the plane that represents the mean orienta- and Hunter, 1982; Rubin, 1987a). These sets and co-sets are them-
tion of the cross strata and the plane that represents the mean selves composed of smaller-scale structures such as wind-ripple,
orientation of the superimposition bounding surfaces defines the grainflow (avalanche), and grainfall strata, the relative proportions
along-crest trend of the superimposed dunes. Assuming that there and distributions of which are determined by the type of bedforms
is no component of along-crest sand transport, then the migration on which the processes responsible for their generation operated.
direction of the superimposed dunes is normal to this trend. Thus, dune elements are larger-scale stratal units composed of a
variety of arrangements of smaller-scale bodies.
Architectural Elements in Eolian Systems Studies of modern dune elements are many and varied.
Downwind changes in dune type across the Great Sand Dunes,
Modern eolian systems are composed of a suite of geomor- Colorado, occur in a predictable sequence from a zone of small,
phic elements that occur on a range of scales, with a variety of spatially isolated, partly cemented dunes at the upwind margin,
geometries and that occupy a variety of positions within the through a zone of undulating barchan, parabolic, and transverse
eolian system. This section illustrates the range of architectural dunes with intervening interdunes, to a zone of large, actively
elements present in modern eolian deserts and discusses the accumulating transverse and star dunes separated by only minor
influences on the development of these elements. interdune depressions at the downwind margin (Andrews, 1981;
Fryberger 1990a). Complex morphological arrangements of star
Dune Elements.— dunes in the Gran Desertio Sand Sea, Mexico, accumulated under
the influence of multidirectional wind regimes through the merg-
Eolian dune elements are accumulations of strata generated by ing of smaller crescentic and reversing dunes (Lancaster et al.,
the migration of hierarchies of eolian bedforms of differing sizes 1987). Internal dune structures in the Al Liwa sand sea, Abu
EOLIAN FACIES MODELS 41

Topset and lee side accretion deposits


A grainfall laminae
grainflow strata
cone-shaped
grainflow foresets
climbing-
ripple strata

wind ripple strata


at dune toe set

Planation surface with plan view geometry of lee slope strata


B grainfall lamination
climbing-ripple
stratification

grainflow (sandflow)
gross stratification

front edge of dune


at time of planation
B

A B 0 6
m

Key
climbing-ripple stratification
dip and strike of
cross strata grainfall lamination
set boundary grainflow (sandflow) lamination

FIG. 22.—A) Schematic diagram showing the small-scale structures of different types of foresets: simple cone-shaped grainflows,
grainfall laminae, and climbing-ripple strata. Plane-bed lamination is often developed on exposed dune crests but is not shown
here. B) Map and cross section of dune foreset cross strata exposed on a planed-off sinuous transverse or barchanoid ridge
dune, showing the distribution of small-scale foreset structures. Simplified from an exposure on Padre Island, Texas, U.S.A.
After Hunter (1977).
42 NIGEL P. MOUNTNEY

A plan viewmain
face of dune
N bounding surface
cross stratum
Transverse dune
trench (b)
side trench section SE
n 4
side tio NW
ec m
trench (c) dir 0
nd
wi
main trench section
SW NE
10
base m
of trenc
0 m 10 h
0
base of dune

main
B plan trench (b) N Barchanoid dune
n
tio
ec

bounding surface
dir

cross stratum
nd

side
wi

trench (c)
main trench section
poor
SW exposure NE
6

poor m
exposure
covered
0
base of dune
side trench section 0 m 8

NW SE

concealed

base of trench
base of dune

C plan
main trench
north wall (b)
N
bounding surface Dome dune
n
ectio cross stratum
win d dir
main trench side wall
south wall (c) (not shown)
0 m 15 dune crest E
main trench, north wall windward (stoss) surface
W

base of trench concealed


main trench, south wall
W small swale on dune surface lee surface
4
E m
0
concealed base of trench concealed

FIG. 23.—The structure of the interior of different dune types revealing various patterns of cross bedding. After McKee (1966) and
McKee (1979). A) Transverse dune. B) Barchanoid ridge dune. Of the modern dunes excavated, this particular example shows
a complexity of internal lamination that would not have been expected from the external morphology and suggests a complex
evolution. C) Dome-shaped dune.
EOLIAN FACIES MODELS 43

A B

C D

0.5 m

FIG. 24.—Examples of ductile soft-sediment deformation structures in eolian sandstones. A, B) Small-scale liquefaction structures in
wind-ripple- and grainflow-dominated eolian dune sandstones. Cedar Mesa Sandstone, Permian, Utah, U.S.A. C) Larger-scale
liquefaction structure affecting several dune sets. Cedar Mesa Sandstone. D) Helsby Sandstone Formation, Triassic, England.

Dhabi are architecturally complex, with nested sets of troughs accumulations are associated with three separate constructional
being generated by a combination of lee-slope reactivation, dune erg-building phases, each partially preserved in a complex mo-
superimposition, and alongslope migration of linear spurs saic. Sand for the construction of more recent eolian bedforms
(Bristow et al., 1996). Remote-sensing techniques (e.g., Breed and was derived partly from the cannibalization of older deposits.
Grow, 1979; Breed et al., 1979; Breed et al., 1987) provide an The current humid (interglacial) climate means that many
opportunity to appreciate the spatial variability of dune and modern eolian dune systems are only partly active and currently
associated elements within eolian systems (Fig. 35). are not accumulating. It is therefore difficult to relate them directly
Quaternary eolian systems provide a link between modern to the widespread eolian accumulations observed in the ancient
active systems and their ancient counterparts preserved in the record (Lancaster, 1998). Instead, comparisons between modern
rock record. The Nebraska Sand Hills are a large, stabilized late dune elements and their preserved counterparts are supplemented
Quaternary dunefield, covering 57,000 km2 (Ahlbrandt and by theoretical models for bedform climbing and accumulation.
Fryberger, 1980). Transverse dune ridges with up to 100 m of
relief and smaller, spatially isolated barchans are separated by Dry Interdune Elements.—
low-lying interdune areas. Eolian erosion in a series of blowouts
and fluvial drainage pathways exposes the interiors of many of Interdune flats and hollows that occur between eolian dunes
the bedforms, and analysis of foreset azimuths suggests uniform are considered dry where the depositional surface exhibits no
migration to the southeast (Ahlbrandt and Fryberger, 1980). The evidence of sedimentation that is influenced or controlled by
coastal Akchar eolian system of Mauritania represents the amal- moisture. Dry interdunes are dominated by wind-ripple strata,
gamation of several separate ergs composed of complex linear though eolian plane beds are also observed and dry interdunes
draa, the accumulation and partial preservation of which have subject to deflation are sometimes characterized by granule and
been controlled by eustatic and climatic variations during the late pebble lags (Hunter, 1977). The absence of near-surface moisture
Pleistocene and Holocene (Kocurek et al., 1991a). Dating by 14C means that the degree to which dry interdunes are colonized by
methods and archeological artifacts indicates that the eolian vegetation is minimal and that animal trackways are restricted.
44 NIGEL P. MOUNTNEY

A B

0.5 m

C D

FIG. 25.—Examples of brittle soft-sediment deformation structures in eolian sandstones. A, B) Slab slide—failure of partly cohesive
sand slabs along a plane of weakness, Askja, Iceland. C) Brittle faulting of grainflow strata. Cedar Mesa Sandstone, Permian, Utah,
U.S.A. D) Brecciated sandstone blocks in the toeset region of a dune set. Helsby Sandstone Formation, Triassic, England.

folds contorted due to fold amplitude decreasing,


loading by sand dune wavelength constant

eolian
dune
preserved erosion surface

0 m 2

FIG. 26.—Schematic illustration of the development of contortion in a siltstone unit as the result of the advance of a large eolian dune
across its surface. Lnagra Formation, Upper Devonian, central Australia. Modified after Collinson (1994).
EOLIAN FACIES MODELS 45

A B

C D

FIG. 27.—Examples of non-eolian physical structures associated with eolian environments. A) Desiccation cracks on an interdune
playa lake surface. Sossusvlei, central Namib Sand Sea. B) Aqueous climbing-ripple strata with preserved form sets. Southern
Namibia. C) Mud curls and rain drop imprints. Southeast Spain. D) Sand-filled desiccation crack in mudstone. Lower Cutler Beds,
Permian, Utah, U.S.A.

The geometry of dry interdune elements, such as those of the is influenced by the presence of moisture (Fig. 38). Damp interdunes
Namib Sand Sea (Figs. 35, 36; Lancaster and Teller, 1988), is are characterized by a range of adhesion structures (Fig. 20) and
largely controlled by the spacing and plan-view shape (morphol- minor salt-precipitation structures (Ahlbrandt and Fryberger, 1981;
ogy) of adjoining dunes. Dry interdunes range from spatially Kocurek, 1981). Near-surface moisture encourages colonization by
isolated hollows, completely surrounded by dunes, to narrow plants and animals, and a variety of root structures, burrows, and
but elongate corridors that extend for several kilometers between surface traces on a variety of scales are common (Fig. 29; Ahlbrandt
rows of transverse or linear dunes, to extensive interdune flats et al., 1978; Loope, 1988; Hasiotis, 2002). The geometry of damp-
completely surrounding isolated bedforms (Figs. 35, 36, 37). The interdune elements can differ substantially from that of dry
width, length, and degree of interconnection of adjacent dry interdunes because the presence of moisture acts to stabilize the
interdune corridors typically decrease from the erg margin sediment surface and restricts the availability of sand within damp
toward the center as surrounding dunes increase in size. The interdunes for eolian transport. Documented examples of water-
degree to which dry interdunes are partitioned is partly influ- table-controlled damp-interdune elements in coastal settings in-
enced by the degree of sinuosity of adjacent dune forms. Straight- clude Padre Island, Texas (Hummel and Kocurek, 1984; Kocurek et
crested bedforms promote the generation of straight, uninter- al., 1992), the Oregon Coast (Hunter et al., 1983), and the Dhahran
rupted interdune corridors, whilst highly sinuous-crested area of Saudi Arabia (Fryberger et al., 1983). Examples from inland
bedforms tend to partition interdune corridors into separate settings include parts of the Namib Desert (Lancaster and Teller,
isolated hollows (Fig. 35B, C, 37). 1988) and White Sands, New Mexico (Simpson and Loope, 1985;
Fryberger et al., 1988).
Damp Interdune Elements.—
Wet (Flooded) Interdune Elements.—
Interdune flats and hollows that occur between eolian dunes
are considered damp where the depositional surface is in contact Interdune flats and hollows that occur between eolian dunes
with the capillary fringe of the water table such that sedimentation are considered wet where the water table rises to or above the level
46 NIGEL P. MOUNTNEY

A B

C D

FIG. 28.—Examples of non-eolian chemical structures associated with eolian environments. A) Laminar calcrete profile, Sossusvlei,
central Namib Sand Sea. B) Sandstone pseudomorphs after desert-rose gypsum. C) Silcrete (chert) developed in an episodically
flooded wet interdune. D) Paleosol with nodular calcrete within an interdune unit. B–D) from the Cedar Mesa Sandstone,
Permian, Utah, U.S.A.

of the depositional surface for protracted periods such that the tions in the water table that reflect changes in the balance between
interdune is continuously or episodically inundated by water ongoing sedimentation, subsidence, and regional climate varia-
(Fig. 35C). Mud-, silt-, and sand-grade sediment may be supplied tion (Kocurek and Havholm, 1993). In some situations flooding
to wet interdunes via eolian processes, or via fluvial processes. may occur where infiltration capacity is exceeded because of a
Additionally, carbonate sediments may also accumulate in long- low-permeability horizon at or beneath the interdune surface
lived ponds (Driese, 1985). Sedimentary structures associated that promotes the generation of a perched water table. Docu-
with such settings include subaqueous current and wave ripples mented examples of permanently and episodically flooded wet-
(sometimes with mud drapes), wavy laminae (Kocurek, 1981), interdune elements in coastal settings include the tidally flooded
contorted bedding (Doe and Dott, 1980), desiccation cracks, back-barrier dune field of Guerrero Negro, Mexico (Fryberger,
raindrop impressions, hard-pan crusts, mud flakes, and mud 1990c; Fryberger et al., 1990), Padre Island, Texas (Hummel and
curls (Fig. 27; Fryberger, 1990b). The continuous or episodic Kocurek, 1984; Schenk, 1990; Kocurek et al., 1992), and the Jafurah
presence of water encourages plant colonization and animal area of Saudi Arabia (Fryberger et al., 1984; Fryberger et al., 1988;
activity, and a wide variety of biogenic structures are common Fryberger, 1990c). Examples from inland settings include the
(Ahlbrandt et al., 1978). The preservation potential of delicate Duero Basin dune field, central Spain (García-Hidalgo et al.,
features such as burrows and animal trackways is often enhanced 2002), dune-dammed paleovalleys in the Nebraska Sand Hills
where such features are emplaced on or in a damp muddy (Loope et al., 1995), and parts of the Skeleton Coast dunefield of
substrate. Interdunes may be flooded for a variety of reasons, northern Namibia (Stanistreet and Stollhofen, 2002).
including ephemeral flash flooding from intra-erg rainfall events,
fluvial inundation from beyond the erg margin, localized and Sand-Sheet Elements.—
temporary water-table elevation associated with extra-erg rain-
fall events (Langford, 1989; Langford and Chan, 1989), seasonal Sand sheets are areas of wind-blown sand that lack high-relief
variations in regional ground-water table, and longer-term varia- bedforms (Fig. 39) but are instead characterized by wind ripples
EOLIAN FACIES MODELS 47

A B

0.25 m

C D

FIG. 29.—Examples of non-eolian biogenic structures associated with eolian environments. A) Reptile trackway on bedding surface.
Cutler Group, Permian, Utah, U.S.A. B) Vertebrate indenter mark in dune-foreset facies. Sherwood Sandstone Group, Triassic,
England. C) Near-surface burrows in damp interdune strata. D) Fossilized root structures (rhizoliths). Cedar Mesa Sandstone,
Permian, Utah, U.S.A.

and sometimes low-relief ridge- and dome-like bedforms called geomorphic elements in cold-climate deserts (Koster, 1988), such
zibar (Nielson and Kocurek, 1986). The development of sand sheets as those in central Iceland (Fig. 39B; Mountney and Russell, 2004).
is controlled by a variety of factors, the most important being the Sand-sheet accumulations rarely exceed 10–20 m in thickness in
supply and availability for transport of a surfeit of sand and modern deserts and often border and/or underlie active ergs,
granules (Breed et al., 1987; Khalaf, 1989; Kocurek and Nielson, thereby acting as stable bases over which large dune forms migrate
1986). Sand-sheet deposits may additionally be characterized by (Fryberger et al., 1979; Khalaf et al., 1984), as is the case for the
plane beds and long-wavelength wind ripples termed megaripples northeastern margin of the Australian ergs (Brookfield, 1970).
(Fig. 8D; Fryberger et al., 1992). The presence of vegetation in some
sand sheets can act to restrict sand movement and dune growth Sabkha and Playa-Lake Elements.—
(Kocurek and Nielson, 1986), as can a shallow water table (Fryberger
et al., 1988). Some sand sheets represent the erosional remnants of Sabkhas are low-relief flats where accumulation occurs wholly
what were previously higher-relief bedforms and therefore reflect or partly as a result of evaporite precipitation (and in some cases
the product of a negative sediment budget and widespread defla- carbonate sedimentation). The term sabkha was originally used
tion. Sand sheets covered with an armoring of granule and pebble exclusively for the description of salt flats in coastal desert
deflation lags are the end product of this deflationary process. settings (Evans et al., 1964; Purser and Evans, 1973) but is now
Individual sand sheets vary in extent from localized sand patches also widely used for the description of inland salt flats, which are
(< 1 km2) to major geomorphic features such as the > 100,000 km2 also termed playa basins. Sabkha sedimentation usually involves
Salima sand sheet of the eastern Sahara (Breed et al., 1987; Maxwell interactions between chemical (precipitate) and eolian processes
and Haynes, 1989). Other large sand sheets in hot desert settings and results in the generation of a variety of wavy and crinkly
have been documented from the Ténéré Desert, Niger (Warren, laminae types that are often disturbed by salt-growth structures
1971), and parts of the Namib Desert (Lancaster, 1985). Although such as teepees. Salt precipitation in sabkhas requires periodic
documented less extensively, sand sheets also form important wetting and subsequent desiccation of the surface and is often
48 NIGEL P. MOUNTNEY

Simple set of cross strata Compound set of cross strata Compound sets of cross strata
(no internal bounding surfaces) (internal bounding surfaces) which together form a coset

I I I
S
R

Coset

Set
Set

Set
R S
R
I I I

Superimposition
surfaces

Supersurface-bounded erg sequences Eolian dune


facies
SS Eolian interdune
Compound
facies
draa Cross strata
system
R Reactivation surface
SS
Simple S Superimposition surface
dune
system I Interdune migration surface
SS Supersurface

FIG. 30.—Models illustrating the geometry of reactivation surfaces, superposition surfaces, interdune migration surfaces, and
supersurfaces in eolian systems. The hierarchical nature of the bounding surfaces, as described by Brookfield (1977), is not always
readily identifiable in the rock record. The surfaces do not necessarily break into universally distinct groups by extent or dip angle.
However, higher-order bounding surfaces always truncate lower-order bounding surfaces. Modified after Kocurek (1991).

interdune migration superimposition reactivation


surface surface surface

cross strata

FIG. 31.—Definition diagram for the hierarchical system for FIG. 32.—Schematic diagram illustrating a co-set of scalloped
describing eolian bounding surfaces in compound-cross- cross strata with internal cyclicity. Two distinct scales of
bedded sands and sandstones, as proposed originally by bounding surface are evident within the co-set. Note how
Brookfield (1977). Interdune migration surfaces arise as a bounding surfaces at the base of the sets pass down dip into
consequence of dune migration. Superimposition surfaces corrugated surfaces. This relationship, which can poten-
represent the migration of superimposed bedforms and/or tially occur at a variety of scales, is indicative of eolian dune
scour pits over a larger parent bedform. Reactivation sur- migration that occurs synchronously with accumulation in
faces represent partial deflation of a bedform lee slope and damp, water-table-controlled interdunes. Based on obser-
arise in response to periodic changes in bedform migration vations from the Jurassic Entrada Sandstone, NE Utah,
direction, steepness, speed, and/or asymmetry. U.S.A. Modified from Crabaugh and Kocurek (1993).
EOLIAN FACIES MODELS 49

Cross-bed azimuth data


A Bedform migration vector
Crestline sinuosity migration vector
Scour pit trend

m
d for n Cre
s
Be rati n
o sinu tline
g o
mi ecti mig osity
i r ratio
d n

B
Tre
rm nd
e dfotion trou of
g
B ra n
g o axe h
mi ecti s
d i r

Downlap

Troughs in the
dip section only

Concordant bedding

FIG. 33.—Examples of common complexities encountered in attempting to relate preserved set architectures to the bedforms
responsible for their generation. A) The migration of 3D (sinuous) bedforms with along-crest-migrating sinuosities preserves
cross strata with mean foreset azimuths that are 35 degrees counterclockwise from the bedform migration direction. B) The
migration of 2D (straight-crested) bedforms that are subject to temporal variations in height generates scour troughs with axes
that trend normal to bedform migration. Models generated using the software of Rubin (1987a).
50 NIGEL P. MOUNTNEY

A B

A
C D

A B C D

FIG. 34.—Schematic illustration of the geometric complexity of trough-cross strata. A) A vertical section oriented transverse to the
trough axis reveals symmetrical cross-stratification planes that are apparently concordant with the trough base. B) A vertical
section oriented oblique to the same trough axis reveals cross-stratification planes that apparently fill the trough asymmetri-
cally and downlap onto its base. This illustrates the problems associated with the measurement of foreset dip azimuths from
core or outcrop for the purposes of establishing paleo–transport direction from trough-shaped cross strata. Modified after
DeCelles et al. (1983).

controlled by subtle water-table changes, which in coastal set- Lancaster, 1999). Sediment supply is the volume of sediment of a
tings may be driven by sea-level change. Modern coastal sabkhas grain size suitable for eolian transport generated per unit time.
are documented from parts of the Bahrain coast (Doornkamp et This sediment may form either a contemporaneous or a time-
al., 1980), Dhahran, Saudi Arabia (Fryberger et al., 1983), and lagged source of material with which to construct an eolian
southern Kuwait (Gunatilaka and Mwango, 1987). system (Kocurek, 1999), and can be derived from a variety of
sources. For example, fluvial, deltaic, and lacustrine terrigenous
Non-Eolian Elements.— sands of the Mojave River and fan delta supply the Kelso Dune
Field, California (Sharp, 1966; Kocurek and Lancaster, 1999),
Eolian sedimentary systems are often intimately associated coastal and shallow marine shelf sands supply the Namib Desert
with a range of other depositional environments, including (Corbett, 1993), evaporitic gypsum from Lake Lucero supplies
ephemeral and perennial fluvial systems, alluvial fans, perma- the White Sands Dune Field, New Mexico (McKee, 1966), and
nent lake bodies, and shoreline and shallow marine systems glacial and volcaniclastic sediments supply the Askja Sand Sheet,
(Chan and Kocurek, 1988; Kocurek et al., 2001). The boundaries Iceland (Mountney and Russell, 2004). Sediment availability is the
between these various depositional environments may be either susceptibility of surface grains to entrainment by the wind
sharp or gradational, and non-eolian elements may sporadi- (Kocurek and Lancaster, 1999) and may be controlled by stabiliz-
cally occur within otherwise exclusively eolian systems. For ing factors such as the presence of vegetation, mud drapes,
example, the eastern portion of the Namib Sand Sea is punctu- coarse-grained lags, an elevated water table, or surface binding
ated by a dune-free corridor, 2–4 km wide and 40 km long, and cementing agents. The extent to which sediment is available
utilized by an ephemeral fluvial system (Fig. 35). for eolian transport determines the actual sediment transport rate
for a given wind. Transport capacity is a measure of the potential
Controls on Eolian Systems sediment-carrying capacity of the wind and increases with wind
power. Where sediment supply and/or availability is limited, the
The preservation of an eolian system in the rock record airflow is unsaturated with respect to its potential sediment load
requires a special set of fortuitous circumstances. Preserved and the wind is potentially erosive. Conversely, an airflow which
architecture reflects the culmination of a series of events that is fully saturated with sediment and which undergoes decelera-
together determine how the geomorphology of the original sys- tion must drop some of its load, thereby encouraging growth of
tem relates to its preserved sedimentary expression. A variety of eolian bedforms (Figs. 40, 41).
external (allogenic) controls determine how eolian systems are
constructed and undergo accumulation, how their accumulation Accumulation of Eolian Successions
is terminated, and how sequences of eolian accumulations can be and Bedform Climbing.—
preserved into the long-term rock record.
Accumulation is the generation of a body of strata by the
Construction of Eolian Systems.— passage of sediment from above to below the accumulation
surface (Kocurek and Havholm, 1993) such that the level of that
Erg construction is a function of sediment supply, sediment surface rises over time. The alternatives to accumulation are
availability, and the transport capacity of the wind (Kocurek and bypass, whereby the level of the accumulation surface remains
EOLIAN FACIES MODELS 51

A
B

C
N
5 km

B C
FIG. 35.—Image of part of the central Namib Desert. A) Separate elements composed of morphologically distinct bedform types are
evident. B) Complex linear draa with superimposed transverse dune ribs. Net sand transport is from SSW to NNE. C) Mosaic of
pyramid star draa with isolated interdune hollows. White color represents salt and calcrete deposits, green color represents
ponded water in wet interdunes. Image courtesy of NASA Earth Observatory collection.
52 NIGEL P. MOUNTNEY

A B

200 m 300 m

C D

1 km 10 km

FIG. 36.—Examples of modern interdunes showing a range of geometric configurations. A) Fully enclosed isolated depression. B)
Fully enclosed large interdune flats. C) Open interdune flat. D) Open interdune flats forming linear corridors with some enclosed
and isolated interdune depressions. A–C) Namib Sand Sea. D) Sahara. Image courtesy of NASA Earth Observatory Collection.

constant, and erosion (deflation), whereby the surface falls. The angle of climb is determined by the ratio between the rate of
accumulation surface is defined by a plane that joins bedform downwind bedform migration and the rate of rise of the accumu-
troughs, and sediment lying above the surface (including lation surface (Fig. 42). For most eolian bedforms, the accumula-
bedforms) is considered to be in transport, whilst sediment tion rate is small compared to the migration rate and the resultant
beneath is considered to have accumulated. Whether accumula- angle of climb is low such that subcritical climbing occurs (Hunter
tion, bypass, or erosion occurs is determined by the net sediment 1977), whereby as bedforms move through space they truncate
budget of the system (Mainguet and Chemin, 1983). Where the upper parts of the preceding bedforms in a train and only the
upstream influx exceeds downstream outflux, a positive net basal part of the bedforms accumulate to form sets of cross-strata
sediment budget exists and accumulation occurs (Fig. 42). By (Fig. 43). Critical climbing occurs where the angle of climb exactly
contrast, neutral budgets and negative budgets result in bypass matches the angle of the stoss slope of the bedforms such that the
and erosion, respectively. For a given sediment influx, accumu- entire bedform is accumulated (Hunter, 1977). Supercritical climb-
lation requires a decrease in the downstream transport rate ing occurs where the angle of climb is greater than the angle of the
and/or a decrease in flow concentration over time (Rubin and stoss slope of the bedforms and both the lee and stoss slope
Hunter, 1982; Kocurek and Havholm, 1993). A downstream deposits accumulate in a manner whereby laminae can be traced
reduction in sediment transport rate occurs with a downwind uninterrupted between successive sets. Both critical and
decrease in wind power (deceleration), often as a consequence of supercritical climbing are usually restricted to ripple-scale
airflow moving into topographic basins or because of a spatial bedforms (Fig. 16).
decrease in the pressure gradient. A temporal decrease in flow
concentration is achieved most easily by a reduction in dune Accumulation in Dry, Wet, and
size, perhaps due to a temporal decrease in wind power (Kocurek, Stabilized Eolian Systems.—
1999).
Accumulation of migrating bedforms occurs as a consequence In dry eolian systems, the water table lies substantially below
of bedform climb with respect to the accumulation surface. The the accumulation surface such that moisture plays no role in
EOLIAN FACIES MODELS 53

A bedform lee slope bedform crestline

bedform stoss slope dominant


wind direction
interdune in all cases

D E

FIG. 37.—Schematic illustration of common bedform configurations in plan view. A) Sinuous-crested transverse bedforms with
crestlines of adjoining bedforms 180° out of phase. Note how the interdune flats form spatially isolated depressions. B) Sinuous-
crested transverse bedforms with crestlines of adjoining bedforms perfectly in phase. Downwind decrease in amplitude of
crestline sinuosity to zero. C) Downwind spatial transition from isolated barchan dunes, through a zone of laterally intercon-
nected barchanoid dune ridges, to low-sinuosity transverse dunes. This pattern is a common configuration at upwind erg
margins. D) Longitudinal dunes that undergo a downwind decrease in crestline sinuosity. Note the resultant increase in the
degree of interconnectivity of the interdune flats. E) Downwind spatial transition from isolated barchan dunes to connected
barchans that are transitional into sinuous-crested linear dune ridges with transverse spurs.

influencing sedimentation, and deposition occurs as a result of In wet eolian systems, the water table or its capillary fringe is in
aerodynamic configuration alone (Kocurek and Havholm, 1993). contact with the accumulation surface such that moisture influ-
Interdune flats in dry eolian systems tend to be areas character- ences sedimentation, and deposition occurs as a result of both
ized by accelerating airflow and are therefore sites of potential aerodynamic configuration and moisture content (Kocurek and
erosion where available sediment is swept up and deposited on Havholm, 1993). Because damp sand is less susceptible to eolian
the lee slopes of nearby bedforms. This results in bedform growth entrainment than dry sand, the availability of sediment for trans-
and expansion at the expense of the interdune flats. Bedforms in port in damp interdunes tends to be restricted. A progressive rise
dry systems tend to commence climbing (i.e., accumulating) only in the relative water table is the fundamental mechanism by
once the interdune flats have been eliminated (Fig. 44). Thus, dry which both dune and interdune deposits accumulate in wet
eolian successions are dominated by cross-bedded dune strata eolian systems (Hummel and Kocurek, 1984; Pulvertaft, 1985),
and tend to lack thick accumulations of dry-interdune strata. the angle of climb being determined by the ratio between the rate
54 NIGEL P. MOUNTNEY

Damp-to-dry interdune transition Contorted damp–wet interdune


A

Gently deformed
wind ripple strata

Climbing damp-interdune element Contorted dune and interdune sets

Wavy but horizontal Direct indicator of Silt and mud drapes


lamination angle of climb & rare wave ripples

Cyclical dune scour into damp interdune Brittle deformation in dune toe sets

Grainflow strata Trace fossils


Wind-ripple strata
Bedform migration
Burrows
Adhesion strata Rootlets 0 Typical scale (m) 1
Contorted lamination

B Interdune deposits Self-healing deformation -


slipface collapse during rainfall
onlap dune slipface Erosional
due to rapid Trace fossils truncation
rise in water table concentrated in dune Adhesion structures of underlying set
toesets and at margins and wavy lamination (subcritical climb)
of interdunes

Water Table
T4
T2/3
T1

Convoluted bedding - Encroaching dune Intertonguing of dune toeset deposits with


deformation toesets indicate damp interdune strata indicates bedform
following rise static water table advance contemporaneous with damp-
in water table interdune sedimentation
T1 - Static water table and encroaching bedform T3 - Static water table and encroaching bedform
T2 - Rapid water table rise and static bedform T4 - Gradual water table rise contemporaneous
with bedform advance (intertonguing)

FIG. 38.—Styles of interaction between eolian dune and interdune elements. A) Examples of common dune–interdune facies
associations. B) Environmental significance of dune–interdune interactions. Modified after Herries (1992).
EOLIAN FACIES MODELS 55

Subaqueous Dry eolian


system

Sabkha
Time Wet eolian
system Lag time

B (-) 0 (+)
Available sediment supply

FIG. 41.—Fields for subaqueous environments, sabkhas, wet eo-


lian, and dry eolian systems as a function of the available
sediment supply over time. At any given time the available
sediment supply is the percentage of the substrate covered by
dry, loose sediment. The initial condition is one of a flat
surface at the level of the capillary fringe of the water table.
After Kocurek and Havholm (1993).

tions, where interdune flats parallel the accumulation surface, the


downwind extent of both dune and interdune units is directly
related to the thickness of the sets that accumulate via the angle
of climb (Fig. 44B).
In stabilizing eolian systems, factors such as vegetation and
FIG. 39.—Morphology of modern sand sheets. A) Skeleton Coast, surface cementation influence sedimentation, and deposition
northern Namibia. Note dune field in far distance. B) Askja, occurs as a result of both aerodynamic configuration and the
central Iceland. degree and type of surface stabilization (Kocurek and Havholm,
1993). The mechanisms that cause eolian systems to become
partly or wholly stabilized are many and varied, and thus the
of relative water-table rise and the rate of downwind migration of causes and styles of accumulation within such systems are also
the bedforms (Fig. 44). Accumulations in wet eolian systems tend highly varied. Vegetation acts to disrupt the primary airflow,
to be characterized by downwind-climbing dune strata sepa- leading to a reduction in sand transport capacity. Precipitation of
rated by damp-interdune units. For simple geometric configura- early diagenetic cements acts to restrict the availability of sedi-
ment for eolian transport and therefore promotes accumulation.

qi Transport qo
Field of
Influx sediment accumulation
ce

saturation level Accumulation


fa

surface a
ur

+Dh
rs

= Sediment
pe

in transport
su

h=0
Actual upwind Field of erosion
ss

Accumulation
transport rate
pa

(Erosional supersurface)
By

s-strata -Dh
Set of cros
Potential upwind
transport rate
Foreset

0
0 1
Downwind deceleration Downwind acceleration
FIG. 42.—Generation and accumulation of sets of cross strata by
Change in transport capacity =
Potential transport rate at downwind boundary
Potential transport rate at upwind boundary
migrating and climbing bedforms, where all transport is
downwind in the section depicted. Accumulation requires a
positive net sediment budget whereby influx (qi) exceeds
FIG. 40.—Fields of accumulation, bypass, and erosion (deflation- outflux (qo) and the height (h) of the accumulation surface rises
ary supersurface generation) defined by the sediment satura- through time. The ratio between the rate of rise of the accumu-
tion level of the influx and the change in transport capacity lation surface and the rate of downwind migration of the
downwind in the system. The slope of the line of bypass is +1. bedforms determines the angle of climb (α) of the accumu-
Modified after Kocurek and Havholm (1993). lated sets of cross strata.
56 NIGEL P. MOUNTNEY

Transport direction
l a
A l

t supersurface
(former
l a depositional
surface)
bypass
B supersurface

C wet interdune
water table at
depositional
surface

a
2l
D
2t

E
subcritical
climb angle
critical climb
angle
supercritical
climb angle

FIG. 43.—Models for bedform climbing. A) Dry eolian system with 100% loose sand cover (positive angle of climb). λ = downwind
bedform wavelength, α = angle of climb. t = preserved set thickness. Note that wavelength can be measured between successive
interdune hollows, along a former depositional surface, or between successive interdune migration surfaces in an alignment
parallel to the former depositional surface. B) Dry eolian system with < 100% loose sand cover—bedforms grow filling
interdunes but angle of climb is zero. C) Wet eolian system with water-table rise—angle of climb controlled by rate of rise of
water table versus rate of bedform migration. D) Angle of climb same as in Part A but bedform wavelength = 2λ, resulting
preserved set thickness = 2t. E) The effects of a progressive temporal reduction in angle of climb on preserved set thickness for
bedforms of a constant wavelength. Note that critical and supercritical angles of climb are usually restricted to ripple-scale
bedforms. Parts A–D Modified in part after Mountney and Howell (2000).
EOLIAN FACIES MODELS 57

UNDERSATURATED METASATURATED SATURATED


ZONE ZONE ZONE
No eolian Positive net sediment budget. Positive net sediment budget.
bedforms Dunes grow at expense of 100% sand cover.
interdunes. Zero climb. Climbing dry erg.

Base level
0% 100%
Increasing sand cover
Theoretical section across upwind erg margin showing zones of
A sand saturation, metasaturation, and undersaturation.

B Accumulation
surface
Climbing
wet erg
C Relative
Water
Table

Capillary
Angle of fringe of Absolute
climb water table Water
Table
A high water table restricts the availability of sediment for eolian transport. As a
result, actual sediment transport rate is typically significantly less than potential Subsidence
sediment transport rate and the wind is not fully saturated.
Eolian bedforms cannot grow to the point where interdune flats are eliminated.
Accumulation in wet eolian systems requires a progressive water-table rise that Relative
protects interdune flats from eolian erosion. Water
The ratio between the rate of water-table rise and the rate of downwind dune Table
migration defines the angle of climb of the system.

FIG. 44.—Models for accumulation in eolian systems. A) In dry systems dune expansion to the point where interdunes flats are
eliminated occurs prior to the onset of accumulation. Based on the discussion of Wilson (1971). B) In wet systems, both dunes and
interdune flats accumulate in response to water-table rise. After Mountney and Jagger (2004). C) Relative water table is
determined by changes in absolute water table and subsidence.

Similarly, the deposition of a thin veneer of mud in episodically ment to another, for example by marine flooding of a coastal erg
flooded interdune areas also acts to retard the effects of eolian system (Glennie and Buller, 1983; Chan and Kocurek, 1988).
deflation (Langford, 1989). Distinguishing supersurfaces in eolian strata from other types
of bounding surfaces is often difficult. Sedimentary features
Supersurfaces and the Generation of Eolian Sequences.— associated with supersurfaces include desiccation cracks and
polygonal fractures (Kocurek and Hunter, 1986), bioturbation,
A cessation of eolian accumulation occurs where the net rhizoliths (Fig. 45B; Loope, 1988), halokinetic (salt) structures,
sediment budget switches from positive to neutral or negative. and large-scale erosional “superscoops” (Blakey, 1988a), all of
Under such circumstances, bedform climbing ceases and is re- which yield important paleoenvironmental information regard-
placed by either bypass (neutral sediment budget) or deflation ing the nature of the accumulation surface at the time of
(negative sediment budget) whereby the wind is unsaturated supersurface formation, but not necessarily about the nature of
such that its potential transport rate is not realized and the the accumulation surface at the time when the eolian system was
accumulation surface falls as the existing accumulation is canni- accumulating. Supersurfaces often have great lateral extent and
balized by the wind. Both bypass and deflation result in the continuity and may bound entire eolian accumulations (Kocurek
generation of a supersurface that caps the underlying accumula- et al., 1991b; Havholm et al., 1993). As such, they are of a higher
tion (Fig. 45; Kocurek, 1988). The accumulation defines a sequence, order than autocyclically generated bounding surfaces and trun-
and its bounding supersurface defines a sequence boundary (Figs. cate all such surfaces. Supersurfaces may be flat lying and planar
46, 47). Deflation occurs until either the net sediment flux once or may exhibit considerable local relief (Kocurek et al., 1991a),
again becomes neutral or positive, or until further deflation is making their recognition problematic. Because most supersurfaces
prevented because the accumulation has been deflated down to, result from external change imposed on a system, their form and
for example, the water table. Deflation in wet eolian systems is any intimately associated deposits may differ from those associ-
controlled by the rate of water-table fall. In either dry or wet ated with other bounding surfaces. Where accumulations of two
eolian systems, where it is the water table that acts to limit the separate eolian sequences are juxtaposed and separated by a
extent of eolian deflation, the resultant supersurface is sometimes supersurface, that surface might be recognized by a change in
called a Stokes surface (Stokes, 1968). Supersurfaces are also gen- the style of sedimentation and/or set architecture on either side
erated as a product of a depositional change from one environ- of the surface. Finally, it may be possible in some situations to
58 NIGEL P. MOUNTNEY

and densely concentrated rhizoliths and thick pebble deflation


A lags may take 104–105 years to form (Loope, 1985). Furthermore,
recent studies of Quaternary eolian systems (e.g., Kocurek et al.,
1991a; Stokes et al., 1997; Swezey et al., 1999; Swezey, 2001;
Bateman et al., 2003, 2004; Bristow et al., 2005) have utilized
techniques such as 14C, thermoluminescence (TL), and optically
stimulated luminescence (OSL) to date separate eolian accumu-
lations. Although timescales vary according to the scale of
supersurface being studied, most workers are now in agreement
that most major supersurfaces represent geologically significant
periods of nondeposition. Indeed, Loope (1985), Clemmensen et
al. (1994), and Rankey (1997) have argued that generation of
supersurfaces in some eolian successions may occur as a result of
Milankovitch-style orbital forcing with periodicities of 18–400
kyr. In many cases, the amount of time represented by eolian
accumulations is likely to be significantly less than the amount of
time represented by the intervening supersurfaces, and, by infer-
ence, preserved eolian successions represent only a small fraction
B of the geologic time over which the eolian systems were active.

Preservation of Eolian Sequences.—


The accumulation of eolian sequences does not automatically
→ ensure their long-term preservation. Indeed, evidence from the
Quaternary record (Kocurek et al., 1991a; Kocurek, 1999) argues
that entire ergs were constructed and accumulated to consider-
able thicknesses, only to be eroded away and leave no record.
Long-term preservation requires that the body of strata be placed
below some regional baseline of erosion, beneath which erosion
does not occur. Thus, the rate of generation of accommodation
and the rate at which eolian accumulations fill that space is a
fundamental control on preserved architectural style (Howell
and Mountney, 1997). The principal agents that generate accom-
modation and promote preservation are subsidence, water-table
rise, sea-level rise, surface stabilization, and exceptional circum-
FIG. 45.—Examples of supersurfaces from modern and ancient stances (Fig. 48).
settings. A) Deflationary supersurface, development of which Subsidence of an accumulation beneath the baseline of ero-
is currently ongoing. Eolian deflation of sand-grade material sion occurs because of tectonism and/or sediment compaction
is concentrating larger clasts at the surface, leading to the (Blakey, 1988b; Blakey et al., 1988). In many systems, the water
development of an armored lag. Huab Basin, northern table defines the baseline of erosion. A rise in the water table may
Namibia. B) Deflationary supersurface separating two sepa- be relative, whereby an accumulation subsides through a static
rate erg sequences. Deflation occurred down to the level of the water table, or absolute, whereby the water table rises in response
water table, thereby enabling plant colonization and the de- to a shift to more humid climatic conditions (Fig. 44C). In the case
velopment of a prominent rhizolith horizon. Cedar Mesa of coastal eolian systems such as Guerrero Negro (Fryberger et al.,
Sandstone, Permian, Utah, U.S.A. 1990), changes in water table may be driven by changes in relative
sea level. Sea-level rise may also promote eolian preservation
where ergs are inundated following marine transgression and
correlate supersurfaces laterally into adjoining non-eolian envi- subsequently buried by marine strata, as described for the Leman
ronments and relate them to basinwide events such as marine erg of the Permian Rotliegend Group, southern North Sea (Glennie
transgressions (Chan and Kocurek, 1988). and Buller, 1983). Surface stabilization, for example by the devel-
Rather like sequence boundaries in marine systems, opment of a vegetative cover associated with a shift to a more
supersurfaces can sometimes be generated as a result of geologi- humid climate, may also increase long-term preservation poten-
cally instantaneous (i.e., isochronous) events such as basin-wide tial. However, the factor that enables the stabilization may be
water-table fall due to regional climate change, or their genera- transient in nature, and, once that is removed the accumulation
tion may be time-transgressive (i.e., diachronous). For example, a may be subject to erosion.
supersurface may be generated by the progressive migration
(translation) of entire erg systems (Porter, 1986). ANCIENT DESERT EOLIAN SYSTEMS
Supersurfaces that bound episodes of eolian accumulation
are diastems that represent a hiatus in deposition. The period of The Nature of the Preserved Eolian Record
time represented by a supersurface varies according to the pa-
rameters responsible for its generation and conceivably may Documented examples of ancient eolian systems exposed in
range from a few days to millions of years. The paucity of tools outcrop range from the description of characteristic facies to
with which to date eolian successions means that determination detailed architectural reconstructions of entire eolian sequences
of the length of time represented by supersurfaces is usually and in-depth discussions regarding their response to various
rather speculative. However, sedimentary features such as large forcing factors that operate over a range of geological timescales.
EOLIAN FACIES MODELS 59

a Erosional Bypass Depositional


(change of
Dry Damp Dry Damp environment)
Falling water
table
Unstabilized

Wet eolian
system
Dry eolian Dry eolian Wet eolian
system system system
(negatively (zero angle (zero angle
climbing of climb) of climb, static
Water table/ water table)
dunes) sabkha flat
Stabilized

Vegetated Vegetated or
relict dunes or cemented
trailing margin dunes

Reg Water table/


sabkha flat

b Level of accumulation surface


Height

Height

Water
Water table level
Time table level
Time

Dune sediment Supersurface


Strata indicating sedimentation on a damp surface

FIG. 46.—Types of supersurfaces. A) Classification of supersurfaces based on net sediment budget (erosional, bypass, depositional),
and the nature of the substrate (dry, damp, stabilized, unstabilized). Downward-pointing arrows indicate erosion of the substrate,
horizontal arrows indicate bypass. Modified after Kocurek and Havholm (1993). B) Example of supersurface generation due to
deflation to the water table (Stokes surface) and an illustration of how the rate of water-table rise can determine the thickness of
a preserved succession. Modified after Havholm and Kocurek (1994).
60 NIGEL P. MOUNTNEY

Deflationary supersurface model Bypass supersurface model


A Dunes migrate
B
Dunes migrate
across former across former
erosion surface. erosion surface.
Zero angle of climb Zero angle of climb

Dunes and interdunes Dunes climb


climb preserving preserving cross-
cross-bedded sets bedded sets
and inclined BS’s and inclined BS’s

Dunes cease climbing Flooding event.


but continue to migrate. Fluvial channel
Net deposition is zero and overbank
(eolian bypass) deposits

Deflation to water table. Continued migration


Surface colonized by of non-climbing
plants and burrowing dunes across
invertebrates flood surface

Renewed eolian Renewed eolian


accumulation accumulation
generates buries bypass
new sequence supersurface

New deflation event New flood event


generates another generates another
deflationary bypass
supersurface supersurface

FIG. 47.—Models for the accumulation and preservation of eolian sequences bounded by deflationary and bypass supersurfaces. A)
Deflationary supersurface model in which eolian accumulation precedes partial deflation to the level of the water table.
Supersurfaces in this example are characterized by rhizoliths, though features such as desiccation polygons, salt structures, and
bioturbation may also be present. Modified after Loope (1985). B) Bypass supersurface model in which eolian accumulation
precedes bypassing due to interdune flooding. Modified after Langford and Chan (1988).

Whilst the Permian to Jurassic eolian outcrops of the Colorado climate change and the generation of stratal bodies and their
Plateau, western United States, have been the focus of most bounding supersurfaces, and a general paucity of material and
intense study in recent years (Blakey et al., 1988), outcrop studies techniques suitable for the absolute dating of stratal units.
from all seven continents now provide a considerable database of
case examples (Table 1). Well studied regions include the Permo- Dry Eolian Systems
Triassic of Europe, the Proterozoic of the Indian subcontinent, the
lower Paleozoic of the central USA, and the Mesozoic of eastern Dry eolian systems, where accumulation is controlled by
South America and southwestern Africa. aerodynamic configuration alone, are widely documented in the
Although studies of ancient systems are blessed with several rock record. The Page Sandstone of Utah and northern Arizona is
advantages over those of modern systems, they also face numer- a dry eolian system represented by separate supersurface-bounded
ous problems. Advantages include an ability to trace and corre- sequences composed of climbing dune strata and an absence of
late key surfaces of environmental significance over large (erg- extensive interdune-flat strata (Kocurek et al., 1991b; Havholm et
scale) distances, an ability to reconstruct eolian set architecture in al., 1993; Havholm and Kocurek, 1994). The eolian succession
3D and observe complex relationships between bounding sur- interfingers with transgressive marine tongues of the neighbor-
faces, an ability to establish the complex 3D preserved geometry ing Carmel Formation, indicating a coastal erg setting (Blakey et
of and style of interaction between eolian and non-eolian archi- al., 1996). The supersurfaces, which are often overlain by sabkha
tectural elements, and an ability to determine how the accumula- and subaqueous strata, have great lateral extent and are charac-
tion of eolian sequences has been controlled by external forcing terized by polygonal fractures (Kocurek and Hunter, 1986) and a
mechanisms such as geotectonic basin evolution. Problems faced corrugated, erosional relief. Eolian-dominated sequences record
by workers investigating ancient eolian systems include an in- episodic, punctuated accumulation, and, rather than being stacked
ability to directly relate preserved sedimentary architecture to vertically, occupy laterally variable paleo-depocenters. Regional
the processes responsible for its generation and the resultant mapping and correlation indicates that dry eolian strata accumu-
requirement for interpretation, an inability to demonstrate un- lated as a consequence of plentiful sand supply from the nearby
equivocally a link between allogenic forcing mechanisms such as shoreline during periods of marine regression, whilst deflation to
EOLIAN FACIES MODELS 61

accumulation above
A preservation space

water table

Dry eolian system, accumulation and


accumulation preservation space
exceeds accumulation by subsidence
preservation space
preservation space by water table

B accumulation and
preservation space
unfilled accumulation
and preservation space
by subsidence by subsidence

Dry eolian system,


water table
accumulation
does not exceed accumulation preservation space
preservation space by water table

C water
table

preservation space
Wet eolian system, by subsidence
accumulation accumulation and
exceeds preservation accumulation preservation space
space by subsidence by water table

unfilled preservation space


D by subsidence

water table
Wet eolian system,
preservation space
accumulation does not by subsidence
exceed preservation accumulation accumulation and preservation
space by subsidence space by water table

FIG. 48.—Components of accumulation and preservation space for dry and wet eolian systems. A) Dry eolian system where the
accumulation has built above the preservation space because a positive net sediment budget exists. Long-term preservation
potential of that part of the accumulation above the preservation space line is low. B) Dry eolian system where the accumulation
has not filled the available preservation space. C) Wet eolian system where water-table rise has enabled the accumulation to
build above that preservation space generated by subsidence. A fall in water table results in deflation. D) Wet eolian system
where the water table is below the preservation space line and the preservation space remains unfilled. Modified after Kocurek
and Havholm (1993).
62 NIGEL P. MOUNTNEY

the water table and supersurface generation occurred during angle of climb in the upper part of the succession reflect progres-
periods of reduced eolian sand supply during marine transgres- sive loss of accumulation space as the erg filled the basin. Eolian
sion (Havholm and Kocurek, 1994). activity ended abruptly as a result of the emplacement of flood
The Jurassic Navajo Sandstone of southwestern Utah (Fig. basalts of the Etendeka igneous province across the erg (Jerram et
49A; Middleton and Blakey, 1983; Herries, 1993) and the Juras- al., 1999; Jerram et al., 2000). These fluid lava flows “drowned”
sic Wingate Sandstone of southeastern Utah (Fig. 49B; the erg, preserving bedforms with heights and downwind wave-
Clemmensen and Blakey, 1989; Clemmensen et al., 1989) are lengths of up to 100 m and 1.3 km, respectively (Mountney et al.,
both examples of predominantly dry eolian systems that were 1999b). Subsequent erosion has exposed the cores of these
subject to episodic fluvial incursions into their erg margins, bedforms and has enabled the original bedform morphology to
possibly as a result of Milankovitch-style climatic fluctuations be related directly and unequivocally to the preserved bounding-
(Clemmensen et al., 1994). surface architecture (Figs. 50, 51). Similar preservation of eolian
The Cretaceous Etjo Formation of northwestern Namibia bedforms following burial by basalt lava flows has also been
represents the deposits of a dry eolian system (Mountney et al., documented from the Cretaceous Botucatu Formation in the
1998, 1999a) and indicates rapid basin infilling by eolian strata Paraná Basin of Brazil (Scherer, 2000, 2002), both episodes relat-
that is characterized by superimposition surfaces that record the ing to the initial onset of breakup of West Gondwana and the
migration of crescentic oblique dunes over larger, slipfaceless opening of the South Atlantic Ocean.
transverse draa (Figs. 50, 51). The preserved thickness of indi- Other examples of preserved but partly denuded and re-
vidual eolian sets varies from 52 m in the basin depocenter to only worked relict dune topography include the Permian Leman
8 m at the basin margin (Mountney and Howell, 2000) as a result Sandstone of the UK southern North Sea (Glennie and Buller,
of increased angles of bedform climb toward the basin center 1983) and the Permian Yellow Sands of northeast England
made possible by the increased availability of accommodation in (Clemmensen, 1989; Chrintz and Clemmensen, 1993), both of
this region. Temporal reductions in preserved set thickness and which are examples of dune flooding by rapid transgression of
the Zechstein Sea. Relict dune topography is also recorded from
parts of the Jurassic Entrada Sandstone in Utah (Eschner and
Kocurek, 1986, 1988) and in New Mexico (Benan and Kocurek,
A 2000), and from the Permian White Rim Sandstone in southeast-
ern Utah (Huntoon and Chan, 1987; Kamola and Huntoon, 1994).

Wet Eolian Systems

Wet eolian systems, in which accumulation is controlled by


progressive water-table rise that occurs in conjunction with ongo-
ing eolian activity, are less widely recognized in the rock record
than dry systems. The Jurassic Entrada Sandstone of the Colo-
rado Plateau (Fig. 52) represents the accumulation of a coastal to
inland eolian system that is composed of a complex arrangement
of eolian dune, damp–wet interdune, and sabkha elements
(Kocurek, 1981). Flat-bedded sabkha deposits immediately over-
lie the uppermost surface of the marine Carmel Formation
(Kocurek, 1981; Crabaugh and Kocurek, 1993; Carr-Crabaugh
and Kocurek, 1998), and dune elements separated by damp and
wet interdune flat elements rise off this basal surface and can be
B traced downwind for several kilometers, where they climb through
the stratigraphy at angles of a few tenths of a degree. In the middle
part of the Entrada succession, climbing dune–interdune ele-
ments are truncated against more flat-bedded sabkha deposits in
a relationship that represents the termination of erg accumula-
tion and the generation of a supersurface that formed as a result
of either a static or falling water table (Crabaugh and Kocurek,
1993). The upper part of the succession is again composed of
downwind-climbing dune and damp or wet interdune elements
that signify accumulation controlled by a progressively rising
water table. Because the sabkha accumulations off which dune
elements climb represent a former depositional surface, the down-
wind distance between the points at which dune and interdune
elements rise off the sabkha surface can be used as a measure of
downwind dune wavelength and interdune-flat width, respec-
tively (Kocurek, 1981). Estimates from the lower part of the
succession suggest that the dune elements were generated by
bedforms with downwind wavelengths of 600–700 m, whilst the
adjacent interdune flats were up to 250 m wide.
FIG. 49.—Examples of dune sets in dry eolian systems. A) Navajo Interactions between water-table-controlled eolian dune and
Sandstone, Jurassic, north of Moab, southeastern Utah, U.S.A. ephemeral fluvial systems have been documented from the
B) Wingate Sandstone, Jurassic, Upper Indian Creek, south- Triassic Wilmslow, Helsby, and Ormskirk Sandstone Forma-
ern Utah, U.S.A. tions, NW England (Øxnevad, 1991; Herries and Cowan, 1997;
EOLIAN FACIES MODELS 63

A B

C D

25 m

FIG. 50.—Eolian dune bedforms and strata from the Etjo Formation, Cretaceous, NW Namibia. A) 52-m-thick cross-stratified set with
asymptotic base. This is the thickest single bed-set documented from anywhere in the world. Person for scale. B) 90-m-high
compound draa drowned by basalt. C) Series of three stacked transverse dune forms, each covered by basalt. D) Barchan dune
draped by basalt.

Mountney and Thompson, 2002). Eolian successions are charac- Stabilized Eolian Systems
terized by lenses of damp and wet interdune strata that exhibit
an intertonguing, transitional relationship with the toesets of Stabilized eolian systems, in which accumulation occurs be-
overlying eolian dune elements, signifying dune migration that cause agents such as vegetation or cementation act to restrict the
was contemporaneous with water-table-controlled accumula- availability of sediment for transport, are not widely recognized
tion in adjacent interdunes (Figs. 53, 54; Pulvertaft, 1985). Down- in the rock record. Part of the Tertiary Tsondab Sandstone, which
wind changes in the geometry and facies of the interdune underlies much of the modern Namib Desert, is characterized by
elements indicate periodic interdune expansion and contrac- sets of cross-stratified dune strata that contain abundant plant
tion in response to changing groundwater-table level and epi- root structures. The prevalence of these structures throughout
sodic flooding (Fig. 55). Sets of cross strata and their bounding much of the succession indicates that the bedforms were veg-
surfaces represent the products of both oblique migration of etated to some degree during their construction, and the accumu-
superimposed dunes over slipfaceless, sinuous-crested parent lation most likely represents an ancient stabilizing dune system
bedforms, and lee-slope reactivation under non-equilibrium (Kocurek et al., 1999). Prior to the development of widespread
flow conditions (Mountney and Thompson, 2002). land-based vegetation in the Devonian, chemical and physical
Some spectacular examples of various styles of eolian–fluvial factors, rather than biogenic factors, were the main stabilizing
interaction are exposed in Pennsylvanian–Permian Lower Cutler agents. A high water table, surface and near-surface cementation
Beds of the Paradox Basin, southeast Utah, U.S.A. (Fig. 56). Stratal by chemical precipitates, periodic flooding, and the presence of
relationships between eolian and fluvial strata document fluvial coarse-grained lags are all considered to enable accumulation in
incursions along open interdune corridors, the ponding of flood warm-climate stabilizing eolian systems (Kocurek and Nielson,
waters by eolian dunes, the incision of large fluvial channels into 1986), whilst permafrost may play an important role in enabling
eolian accumulations, and the reworking of eolian dune and accumulation in cold-climate systems (Mountney and Russell,
interdune deposits by intra-erg fluvial systems. 2004). Sand-sheet deposits in the Upper Silurian Tumblagooda
64 NIGEL P. MOUNTNEY

FIG. 51.—Depositional models illustrating the temporal evolution of the Etjo Sandstone Formation. A) Restricted sediment supply,
underfilled basin. B) High sediment supply, rapid infilling of accumulation space. C) Exceptional preservation of original
bedform morphologies because of flood basalt emplacement. Modified after Jerram et al. (1999) and Jerram et al. (2000).
EOLIAN FACIES MODELS 65

moisture, possibly fog, and minor salt deposition acted as the


binding agents (Trewin, 1993). Early Proterozoic eolianites of the
Whitworth Formation, Queensland, Australia, are characterized
by cross-stratified sets with irregular, overhanging tops and
pseudomorphs of gypsum and anhydrite (Simpson and Eriksson,
1993). These deposits are considered to have accumulated as a
result of early cementation by evaporites and/or adhesion of
damp sand within the capillary fringe. Despite an abundant
sediment supply and a net aggradational setting, sand-sheet
deposits of the Proterozoic Shikaoda Formation, India, were
stabilized to the point where dunes did not develop but rather
sand-sheet accumulation occurred, short-term preservation of
which was enabled by repeated flooding into the low-lying
supratidal flat region, whilst long-term preservation was con-
trolled by subsidence of the accumulation below the water table
(Chakraborty and Chakraborty, 2001).

FIG. 52.—The Entrada Sandstone, Jurassic, exposed in Arches Mixed Eolian Systems
National Park, southeast Utah, U.S.A.
Some eolian successions exhibit features characteristic of dry,
wet, and stabilized systems. The Permian Cedar Mesa Sandstone
Sandstone of western Australia accumulated through surface of southeast Utah, U.S.A. (Loope, 1984) exhibits a complex spatial
binding or cementation, as indicated by the presence of corru- variation in sedimentary architecture which, in terms of paleo-
gated tops to dune sets, broken beds, wind-deflation ridges, and geographic setting, reflects a transition from a dry erg center,
open burrows, all of which suggest that a combination of surface through a water-table-controlled eolian-dominated erg margin,

0 2 4 6 8 10

meters

Facies Sedimentary structures


B C Vegetation Cross-bedding Sand volcano

Eolian dune Contorted


Flame structure
foreset bedding
Eolian dune Nearly horizontal Rip-up
toeset wind ripple pebble clast
Eolian dune
Wave ripple Desiccation crack
flank
Footprint indenter
Dry interdune Wavy lamination mark and burrows

1m 1m Damp interdune Planar lamination Massive

FIG. 53.—Eolian dune–interdune architecture in the Wilmslow Sandstone Formation (Sherwood Sandstone Group), Triassic,
England. A) Dune sets with numerous reactivation surfaces are interbedded with damp interdune units characterized by wavy
lamination. B) Detail of dune set and underlying wavy-laminated interdune. C) The top of the formation is delineated by a fluvial
incision surface.
66 NIGEL P. MOUNTNEY

o
175 > <0 0 5 10 15 20 25 Facies Sedimentary structures Symbols
13 14 15 7 5o
12 1
10 11 Cross-bed
o
8 9 2 Fluvial Cross-bedding Sand volcano 20>224
<355 7 dip azimuth

1 2 3 4 5 6 3
4 meters Contorted Bounding surface
5 Fluvial (vf sand) Flame structure 12>162
bedding dip azimuth
20
6 No vertical exaggeration 180 >18
o
19 <0
Rip-up Translatent
19 7 16 17 1 85o Eolian dune Wind ripple TS
14 15 pebble clast strata
8 13 2
18 25 11 12
9 5o> 8 9 10 3
Dry interdune Wave ripple Desiccation crack GF
Grainflow
10 o
6 7 4 strata
17 11 <000 3 4 5
5
2
12 1
6 Set identifier
16 26 o Damp interdune Wavy lamination Indenter mark H06
7 5> label
15 20 F17 8 o

9 Wet interdune Planar lamination Massive


Modern 175 >
14 19 vegetation 29 30
F06
27 28

meters
13>310 25 26 20
13 18>305 16>336 24
18 F05 06>080 23
F07
21 22
19 20 19
12 17 18>315 18
17>310 F16
16 17
F15
F04
14 15 18
meters

11 16 F09 12 13
F03 10 11
o
8 9 17
10 15 <355 5 6 7
F04
3 4
F10
G0 2 16
9 14 F09 F08 F10 8 1
D06

16>288
15
8 13
TS
16>280
F18 20
04>268 F10
F10 14
7 D03 12 F01
08>220 24>250 F13 19
22
>2 13
6 D05 64 11 F13

meters
D02 D07 16 10
>2
52 >2
38
18
18>230 E01 18 12
5 TS 14>240
16>235 26>350 16
>2
36 10 G0
7
D04 17

meters
>2
24>326 20 54 H12
10>220 >2 11
4 44 22 TS
9 F13
F12
16>305 D01
>3
44 16
F11 10
3 8 H17
F02 15
F12
G0 9
2 24
>2 7 6

20
40 14 H 22
>2 TS 8
15 46 6
1 14 08>080 F12

12 13 1 E02
24>240 13 H16
11 20>224 G0 H20
18>264
10 18 03
E
12>162 24>012 20>230
5
7
0 8 9 2 >2
54
5 F14 H21 H19 H15 GF TS
7 20>222 25>012 12
6 3 21>210 20
5
meters 19>030 >2 G0 H18
4 TS 24 4 6
3 4 4
1 2 11
H07

meters
20 H13
5 >2
24 5
3
16>244
me 6
18 19 G0 G0
10 12>220
ter 7 17 3 9 H06 4
2 16 1 H09
s 14 15 G0
9 14>280 H08
E

8 13 2 20 3 H14 H05
11 12 >2
24 G0 TS 3
1 10 4
S

9 9 3 G0
TS
8
meters
7 8 1 20>290 16>310
<0 10 6 4 H14 H12
2
5
7 101 4 H07 H03
20 1 5o 12 1 2 3 me5 G0
2
7
16>254
14>230 08>310 H03
H02 Plaque

ter 6 1
N

H10 H11 H01


H04
2
19 s 7 6 10>230 H06 H02
14>300
W

3 14>308 0
o <0 8 12>308

160 >
18 4 7 5
5 1 0o 30
19 20 10 28 29
17 6 18 2 4 27
16 17 25 26
7 14 15 3 23 24
16 12 13 3 21 22
8 11 4 20
9 10 19
o 5 18
15 9 25 o <340 6 7 8 2
15 16 17
5>
meters
10 4 5 6 14
3 12 13
14 11 1 2 7 1 10 11
8 9
12 8 7
5 6
13 9 0 4
2 3
10 1
12 11 o
19
175 > <065o
12 o
25 o <355
meters

11 6 7 8
18 13 0> 5 9 10 o
175 >
18>310 3 4 11 12
10 14 1 2 13 14 15 o
17 8
15 16 17 18
245o> <355 6 7
12>300 19 4 5
9 20 3 20
16 21 22 1 2
23
8 20>300 14>250 19
15 19
12>240
7
20 18
14 18
24>348 19
6 17
13 17
5
I09
18 16
I08 J08
12 16
I09

4 17 15
J11
11 15
meters

J09
3 16 14
I10 10 12>280
14
2 I04 16>255 15 13
M05
9 J08
13
I07 18
12>250
I07
1 18>312 >2
56
14 12
08
>2 8 18>350
J07
18>285 22 12
>2
I09 55 14>276 18>270 20>248

meters
30
0 K01 M02
13 11
14 7 20
>2 10>265 >2 11
12>335
meters

95 12
TS 14 14>232 TS
> 26 08>228 12 10
14 I0
>3 3
24 24>196 M04
I05 65 J10
>2
38 10 L01 22>252
0 10
I01
12>282 12>290 J10 GF
K06
TS 20>246 12>230 11

meters
1 12>274 10>250 9
20 14
5 18>288 TS J06
>3
00
>3
30
22
>2 9 GF
2 18>320
J05 28
GF 24>2
GF
10
I02 I06 12>290 10>196 10>226
14
I05
3 14>258 M03 8
3 >3 4 J04 6 K03
8 18>284
14>276
10>200
00 J03 TS 20>300
4 J02
K0
7 L02 9 16>198 N02
08>250 18>284 7
3 J01
5 1 7 20>296 12>234
19 20 20
>2
20
>2
6 18 2 66 76
08>232 8 6
17 K02
me 7 2
14 15 16 3 TS 6 12>238 18>220 10>23
ter 11 12 13
4
14
>2
12
>2
26
>2 7 14>282 5
s 8 1
9 10 80 70
K05 18
30
5 M01
18>228
14>250
GF N01

meters
9 8 5 >2
7 20>6350
TS 56
5 6 12>250
4
10 0 4 6 4
2 3 10
>2 10>256
11 1 7 36 5 3
12 me 8 K04 14 3
>2
24 4
ter 9 2
s 10
2
3 1
11
1
12 2
6 7 8 0
13 0 5 9 10
3 4 11 12
14 1 2 13 1
14 15 8
15
16
meters16 17 18 19 0 4 5 6 7
20 3
meters
21 22 1 2
23

FIG. 54.—Sedimentary architecture of part of the Helsby Sandstone Formation (Sherwood Sandstone Group), Triassic, England.
Modified after Mountney and Thompson (2002).

to an outer erg margin subject to episodic fluvial incursion and table, and regionally extensive deflationary supersurfaces devel-
stabilization (Langford and Chan, 1988, 1989). Accumulation in oped during periods of widespread water-table fall (Fig. 47;
the erg margin was controlled by periodic water-table rise coupled Loope, 1985; Langford and Chan, 1988). A spectrum of deposi-
with ongoing dune migration and associated changes in the tional models that reflect a range of dry to wet and partly
supply and availability of sediment. Variation in the level of the stabilized eolian elements is envisaged to account for the complex
water table relative to the accumulation surface determined the architecture of the succession (Figs. 58, 59).
nature of interdune sedimentary processes, and a range of dry,
damp, and wet (flooded) interdune elements are recognized CONSTRUCTION OF FACIES MODELS
(Figs. 57, 58; Mountney and Jagger, 2004). Variations in the FOR DESERT EOLIAN SYSTEMS
geometry of these units reflect the original morphology and
migratory behavior of dunes and spatially isolated dry interdune Approach to Facies Modeling
hollows in the erg center, locally interconnected damp and/or
wet interdune ponds in the eolian-dominated erg margin, and Studies of modern and ancient eolian systems document a
fully interconnected, fluvially flooded interdune corridors in the wide range of facies variability and architectural complexity and
outer erg margin. Relationships between eolian dune and demonstrate that there is no such thing as a “typical” eolian
interdune units indicate that dry, damp, and wet interdune system. As a consequence, it would be counterproductive to
sedimentation occurred synchronously with eolian bedform mi- attempt to explain eolian systems in terms of a single facies
gration. Temporal variation in the rates of water-table rise and model, nor is it feasible to propose a large number of separate,
bedform migration determined the angle of climb of the erg- highly specialized models that could together account for all
margin succession, such that accumulation rates increased dur- possible variations; this would defeat one of the fundamental
ing periods of rapidly rising water table, whilst sediment bypass- purposes of a facies model, namely to act as a norm with which to
ing (zero angle of climb) occurred in the aftermath of flood events compare individual examples. Rather, eolian sedimentology and
in response to periods of elevated but temporarily static water stratigraphy has now progressed to a state where it is possible to
EOLIAN FACIES MODELS 67

I Interdune migration surfaces


03 30
S Superimposition surfaces
0 0
o

12 2
0 10
o
R Reactivation surfaces
o

Superimposed bedforms Slipfaced primary

o
migrate toward 270° bedform
Spatially isolated
Damp–wet Primary bedforms
wet interdune
interdune corridor migrate toward 300°
Minor
Water fluvial
table I channel
S I
Laterally
interconnected I I I
damp interdune I I
strata S S S
R I
I S R R
S I
I I S
n
uccessio
Fluvial s
Isolated damp interdune
strata in trough base
Large-scale disorganized Interdune strata thicken and
4 trough cross-bedding pinch as they climb downwind
Downwind climb in response to water-table changes
m Abandoned fluvial of eolian sets
0 0 m 20 topography with relative to top of
mud drapes fluvial unit

FIG. 55.—Depositional model for the Triassic Helsby Sandstone Formation, Cheshire Basin, UK. Dune elements climb downwind, as
do adjacent interdune elements, which exhibit downwind facies variability that reflects subtle changes in the level of the water
table relative to the accumulation surface during accumulation. Lateral connectivity of the interdune elements is controlled by
dune morphology, and both isolated interdune hollows (ponds) and interconnected, throughgoing corridors are recognized.
Modified after Mountney and Thompson (2002).

construct “tailor-made” facies models that can be designed to tudinal dune models, are static in that they do not fully incorpo-
help solve specific problems. This new approach to eolian facies rate the effects of changes in parameters such as sediment flux
modeling requires two basic types of information: (1) a database and as a consequence do not explain temporal changes in dune
containing a description of the physical attributes of the elements size, angle of climb, and the generation of supersurfaces (Fig.
that are known to make up eolian systems, and (2) a set of rules (i.e., 60). Static models nevertheless remain useful because they act as
first principles) that dictate the various permissible ways in a basis for the fundamental description of new examples, they
which these elements can fit together in order to construct a viable can be used for prediction, they can be used to infer processes of
model, based on our understanding of how physical processes deposition, and they act as a norm against which individual
operate within modern systems. Architectural elements are the examples can be compared. However, such models usually
basic physical building blocks of eolian systems and include cannot explain the complex architectures and juxtaposition of
facies, bed-sets, and bounding surfaces, whilst the physical at- elements that arise as a consequence of changes in allogenic
tributes of these elements are descriptions of parameters such as controlling parameters.
likely size and shape. Successful facies models need to encapsu- Dynamic facies models incorporate the effects of spatial and
late architectural complexity on a range of different scales from temporal variations in one or more of the fundamental control-
the grain and lamina scale, through the bed-set and co-set scale, ling parameters. Key controlling parameters that are known to
the architectural-element scale, right up to the sequence and change both spatially and temporally include the rate of sand
basin scale. supply, the availability of sediment for transport, and the carry-
ing capacity of the wind, which together define the sediment state
Static Versus Dynamic Facies Models of an eolian system and govern its construction (Kocurek and
Lancaster, 1999), the level of the water table and its rate of change,
In static facies models, the various parameters that control the level of the equilibrium height (below which a positive net
the construction, accumulation, and preservation of the eolian sediment budget exists), and the distribution of stabilizing fac-
system remain constant through both time and space and, as a tors, which together determine the mechanism, rate, and style of
consequence, the arrangement of architectural elements within eolian accumulation (Kocurek and Havholm, 1993), and the rate
the resultant models is relatively simplistic. Many early eolian of subsidence, which partly defines the extent to which eolian
facies models, such as erg margin, transverse dune, and longi- accumulations are preserved.
68 NIGEL P. MOUNTNEY

3m

B C

0.5 m

FIG. 56.—Examples of eolian–fluvial interaction in the Lower Cutler Beds, Pennsylvanian–Permian, Utah, U.S.A. A) Fluvial strata
infilling interdune hollow and onlapping onto flanks of eolian dune form. B) Large fluvial channel cutting into older eolian dune
sequence. C) Erosionally based fluvial channel and paleosol between two eolian dune sequences.

Static Eolian Facies Models bedform migration (Mountney and Thompson, 2002). For dunes
and interdunes of a given size, steeper angles of climb (induced
A range of static models can be used to illustrate the styles of either by slow bedform migration or rapid water-table rise)
eolian architecture that can develop depending on sediment preserve thicker dune and interdune units.
availability (a parameter that is often determined largely by the
level of the water table relative to the accumulation surface) and Dynamic Eolian Facies Models
angle of climb of the system (Fig. 61). A continuum from wet to Subject to Temporal Variability
dry systems occurs with increasing sediment availability (Kocurek
and Havholm, 1993). A high water table restricts sediment avail- Figure 62 depicts an example of a dynamic depositional
ability and enables the generation of extensive interdune flats model that incorporates an element of temporal variation. Tem-
separated by eolian dunes of restricted size. Damp interdunes poral variation in sediment availability, which for wet eolian
develop where it is the capillary fringe of the water table that systems may be determined by changes in the level of the water
restricts sediment availability, whereas wet (flooded) interdunes table relative to the accumulation surface, is a fundamental
occur where the water table rises above the level of the interdune control on preserved architecture in many eolian systems, includ-
floor. A low water table increases sediment availability and ing the Cedar Mesa Sandstone (Langford and Chan, 1988, 1989;
promotes eolian dune growth through cannibalization of adja- Mountney and Jagger, 2004), the Navajo Sandstone (Herries,
cent interdunes (Wilson, 1971, 1973), the conclusion of this pro- 1993), the Ormskirk Formation (Herries and Cowan, 1997), and
cess being eolian dune growth to the point where interdune flats the Helsby Sandstone Formation (Mountney and Thompson,
are reduced to isolated dry interdune hollows, as for a dry eolian 2002). Periodic dune expansion and contraction is inversely re-
system (Fig. 61A). A second control on preserved sedimentary lated to interdune extent and chiefly reflects temporal variations
architecture is the angle of climb of dune–interdune units, which in sediment availability and relative water-table level. Changes in
for wet eolian systems (Fig. 61B) is determined by the ratio these controlling parameters could result from a variety of driv-
between the rate of water-table rise and the rate of downwind ers operating over various timescales ranging from seasonal
EOLIAN FACIES MODELS 69

A B

C D

FIG. 57.—Examples of facies architecture in a mixed eolian system, the Cedar Mesa Sandstone, Permian, Utah, U.S.A. A) Single 20-
m-thick erg sequence bounded by deflationary supersurfaces with rhizoliths. B) Small-scale trough-cross-bedded eolian dune
sets. C) Wet interdune unit intertonguing with eolian dune sets and demonstrating coeval existence. D) Fluvial flood surfaces.

flood events to long-term climatic changes induced by regional or relative water table. Note how the resultant architecture varies
even global climate change. The amplitude of the change in from that depicted in Figures 62 and 63A, which ignore the effects
sediment availability and/or water-table level determines the of periodic changes in the angle of climb.
maximum and minimum size of the dunes and interdune flats,
whilst the periodicity of the change in these parameters deter- Dynamic Eolian Facies Models Subject to Spatial Variability
mines the rate of change of dune–interdune size (Fig. 62).
Additional complexity needs to be incorporated where tem- Many erg successions are characterized by interdune ele-
poral changes in certain controlling parameters have an influence ments that progressively increase in size at the expense of eolian
on others. For example, evidence from a variety of outcrop dune elements from the erg center to the erg margin (Porter,
studies (e.g., Mountney and Jagger, 2004) indicates that the angle 1986). This primarily reflects spatial changes in the sediment
of climb of wet eolian systems is influenced by temporal changes saturation level of the wind, which, in downwind erg margins, is
in sediment availability and relative water table (Fig. 63). Whilst often unsaturated with respect to its potential sand-carrying
rapid rates of water-table rise promote interdune expansion, they capacity because of limited sediment supply and/or availability
also act to increase the angle of climb in the system. Conversely, (Kocurek and Lancaster, 1999). Many modern ergs and ancient
slow rates of water-table rise act to reduce the angle of climb, erg systems are characterized by a progressive shift from a dry
whilst a temporarily static water table would induce bypassing eolian erg center, through a damp, water-table-controlled erg
and promote the generation of a bypass supersurface. Figure 63B margin, to a wet or periodically fluvially flooded outer erg
presents a dynamic facies model that incorporates cyclical tem- margin, and thus represent a spatial transition from a dry to a wet
poral changes in both water-table level and angle of climb. Damp eolian system (Mountney and Jagger, 2004). The model in Figure
and/or wet interdune expansion and steeper angles of climb 64 depicts both temporal changes in controlling parameters as
occur synchronously in response to relative water-table rise, water-table level, sediment availability, and angle of climb un-
whilst damp and/or wet interdune interconnection and non- dergo cyclical variation, and spatial changes in controlling pa-
climbing bypass occur synchronously with a high but static rameters from an erg-center to an erg-margin setting. The 2D
70 NIGEL P. MOUNTNEY

Predominant fluvial flow


Prevailing wind direction - SE Predominant Fluvial Flow
Prevailing wind direction - SE - W to SW
Source-bounded fluvially - W to SW Passive flooding of
A entrenched eolian dune
Overbank and
interdune area
B
Braid channel sheetflood deposition
away from main channel

Su
Flu pers R

via urfa Flu


lly
do ce - d
R

via
mi lly
na eflat Flood carrying do
mi
ted ion na
erg to g
R

extrabasinal sediment ted


ma rou and reworking eolian erg
rgi nd R

ma
n wa sediment rgi
ter n
tab
le R

Eo
lian Eo
-do lian
min R R

-do
ate R R R
mi
de na
SS rg R
R

ted
ma erg
R

rgi ma
n
R

Freshwater limestone rgi


R

n
pond, bioturbated prior
to fluvial deposition
Major fluvial flood, overlying a rooted Fluvially reworked eolian sandstone,
deflationary supersurface. Edge of onlapping on to eolian dune.
entrenched eolian dune, with fluvially Generated as flood breaks into
reworked top active dune field

R
R

Very thin, laterally restricted fluvial


SS SS
facies in the outer erg.

SS
Small-scale eolian and fluvial
Outcrop exposure showing two
interaction in the eolian-dominated
major fluvial incursions over Elliptical mud-filled interdune
erg margin, along a flooded
deflationary supersurfaces at the element , with groundwater calcrete
interdune corridor
fluvially dominated erg margin developed at deflationary super-
surface. Eolian erg margin

FIG. 58.—Facies models of eolian and fluvial interactions in the Permian Cedar Mesa Sandstone, SE Utah, U.S.A., from the fluvial-
dominated erg margin to the eolian-dominated erg margin. A) Major fluvial incursions over a low-relief, deflationary
supersurface following erg deflation to the level of the ground-water table. B) Styles of eolian–fluvial interactions during
period of active erg accumulation. After Jagger (2003).

model, which for simplicity depicts the migration of simple Rather, the encapsulation of 4D complexities is perhaps best
transverse bedforms, portrays spatial variations in an orientation undertaken by building a purpose-designed facies model from a
perpendicular to the trend of eolian bedform crestlines. How- set of simpler constituent parts (Fig. 65).
ever, most eolian systems also exhibit architectural variability
and complexity in orientations parallel to the crestline trend, and APPLICATION OF A DYNAMIC EOLIAN
an appreciation of the nature and causes of this complexity is FACIES MODEL: AN EXAMPLE FROM
essential. In such sections, architectural complexity is largely THE UK SOUTHERN NORTH SEA
dependent on the morphology of the eolian dunes that populate
the system. The planform wavelength and amplitude of along- Facies models have long been used as an aid in hydrocarbon
crest sinuosities and their rate of along-crest migration (Rubin exploration, where they act as a tool for subsurface prediction
1987a) determine the degree to which interdune corridors be- when employed in conjunction with a limited amount of primary
come interlinked as dunes and interdunes migrate and undergo data from core (Fig. 66), wireline log, and/or seismic. When used
expansion and contraction in relation to changes in parameters for subsurface reservoir prediction and characterization, a facies
such as sediment availability and water-table level. model needs to be sufficiently generic to be applicable at an inter-
well or even basin scale but also needs to be specific enough to
Encapsulating 4D Complexity in Eolian Facies Models depict the likely subsurface architectural complexity of a particu-
lar geologic setting. Employment of a valid model can lead to an
From the above discussion it is evident that most eolian improved understanding of the subsurface architecture and can
systems are geometrically and architecturally complex, and for provide insight regarding the range of controls that influenced
facies models to reflect this complexity they must incorporate a the accumulation and preservation of the original sedimentary
large amount of detail on a variety of scales. Models need to system.
account for potential variability in three spatial dimensions plus The Lower Permian Rotliegend Group, which forms the main
the temporal dimension. Given the relatively large number of gas reservoir unit in the UK southern North Sea (SNS), accumu-
controlling parameters and their wide range of permissible val- lated in an intermontane basin and forms a semiarid continental
ues, the notion of a single, all-encompassing model is unrealistic. succession composed of a series of facies belts that pass from
EOLIAN FACIES MODELS 71

D Eolian dune
Wet interdune
0.05–2km Flood surfaces

R R

10–15 km

Cedar Mesa Sandstone Cutler Formation


A 4–40 m

Eolian dune Fluvial channel & overbank interdune

Flood surfaces

C Eolian dune B 50–300 m


Overbank

5–9 m
interdune Fluvial channel
Flood surfaces
Overbank interdune
R

Eolian dune
Rooted horizon R R

FIG. 59.—Schematic model of the Cutler–Cedar Mesa fluvial–eolian interaction indicating the geometry and scale of the intertonguing
deposits. A) Overall geometry of the erg-margin system and relationship of bypass (flood) supersurfaces to fluvial and eolian
strata. B) Geometry and facies relationships associated with fluvial-channel emplacement into eolian dune succession. C)
Relationships between overbank–interdune deposits, eolian dune deposits, and bypass (flood) supersurfaces. D) Relationships
between wet interdune deposits and bypass (flood) supersurfaces. Note the vertical exaggeration and different scales. After
Langford and Chan (1989).

marginal alluvial systems, through erg and erg-margin deposits, indicates that the climate change that acted as a forcing param-
to a sabkha and playa-lake system in the basin center (Glennie, eter during Permian times is likely to have been regional in
1972, 1990). Whilst large gas discoveries of the 1960s (e.g., Leman) extent, and, as a consequence, individual cycles and their bound-
were located in erg-center accumulations, more recent discover- aries are considered to have chronostratigraphic significance.
ies have largely been restricted to “feather-edge” settings where Correlation of cycles between different, coeval depositional
erg-margin accumulations interfinger with alluvial and playa- environments (e.g., alluvial fan, erg, playa lake) led to the
lake facies. Throughout the 1980s and early 1990s, a paucity of age recognition of various drying- and wetting-upward motifs, and
indicators, together with a crude lithostratigraphic framework, has provided a means of correlation from basin-margin alluvial-
meant that prediction of and correlation between feather-edge fan successions, through erg-center dune successions, to a ba-
eolian accumulations with good reservoir potential was prob- sin-center playa-lake succession. Despite maximum accumula-
lematic. tion rates typically being an order of magnitude slower in playa-
Analysis of repeating facies associations observed in a num- lake environments than in eolian-dune and some alluvial-fan
ber of SNS cores led Howell (1992) and George and Berry (1993, systems, individual cycles exhibit similar thicknesses between
1997) to suggest that accumulation and preservation of much of these sub-environments because preserved cycle thickness is
the Rotliegend Group may have been externally controlled by considered to be controlled by the rate of creation of accumula-
climatic cyclicity, and that the eolian succession might be di- tion space (Fig. 68). For example, high accumulation rates in the
vided into a number of separate erg sequences, each bounded by eolian part of the system resulted in rapid infilling of the
a regionally extensive supersurface. In an analysis of 55 wells available space prior to the generation of a bypass supersurface.
from the SNS region, Howell and Mountney (1997) recognized By contrast, low accumulation rates in the playa-lake part of the
12 repeating facies cycles across the basin, each characterized by system meant that the available space was filled more slowly
distinctive facies successions indicative of progressive drying- and bypassing did not occur until much later in the climatic
upward then wetting-upward trends that culminated in a cycle (if at all). Thus, whilst the internal architectures of sedi-
supersurface or its correlative conformity at the cycle bound- mentary cycles are controlled by the depositional processes and
aries, which were interpreted to represent points of minimum their positions within the basin, the thickness of the cycles
aridity (Fig. 67). This suggests a regional climatic control on directly reflects the rate of accommodation creation (Howell,
sequence generation and preservation. Paleoclimatic evidence 1992; Howell and Mountney, 1997). The subdivision of the
72 NIGEL P. MOUNTNEY

Eolian facies model — Dry eolian erg center, transverse bedforms

Complex draa with transverse elements Isolated dry interdune hollow Barchan dune migrating over sandsheet
Dune slipface Transverse Barchan Wind-
Dunes superimposed grainflow and dune dune rippled
on slipfaceless draa Dry grainfall sand
interdune
Prevailing wind

Superimposition
surface Adhesion
structures
Sand-
sheet Capillary
fringe of
water table
Interdune
migration
surface

All photo examples


Dry Dune top- Dune core Dune plinth Eolian Damp–wet from Central Namib
interdune set wind- grainflow & grainflow & sandsheet interdune
wind-ripple ripple grainfall wind ripple wind-ripple adhesion Desert
strata strata strata strata strata strata

Eolian facies model — Periodically flooded erg-margin system

Desiccation cracks, lake margin Dead Vlei — a dry desert lake from the air Sossusvlei — a seasonal dune-dammed lake

Lake Lake-margin Sheetflood


sabkha deposits Sandsheet
with small
eolian bedforms
Dune-dammed
lake with
evaporitic fringe

Prevailing wind

All photo examples


from Central Namib
Desert
Playa Lake-margin Inland Ephemeral Ephemeral
lake sabkha sabkha lake margin lake
Damp–wet Lacustrine
interdune mudstones
strata

FIG. 60.—Simple facies models for eolian systems. A) Model for the accumulation of simple and compound transverse bedforms in
a dry-erg-center setting. B) Model for the accumulation of siliciclastic sabkha flats and isolated eolian dunes in an episodically
flooded erg-margin system. Modified after Howell (1992).
EOLIAN FACIES MODELS 73

A Extent of interdune flats B Angle of climb


Dry eolian system
Increasing sediment supply and availability

Increasing angle of eolian bedform climb


No interdune flats (only dry interdune hollows) Critical angle of climb (rarely realized in nature)

Interdune area < Bedform area Subcritical angle of climb

Interdune area = Bedform area Subcritical angle of climb

Wet eolian system

Interdune area > Bedform area Zero angle of climb

Extent of damp–wet interdune flats increases as Angle of climb in wet eolian systems controlled
eolian sediment supply and its availability for by ratio between rate of water-table rise
transport decreases in response to rising water table and rate of downwind bedform migration

FIG. 61.—Basic controls on interdune geometry in wet eolian systems. Modified in part from Kocurek and Havholm (1993).

upper part of the Rotliegend Group into separate sequences intrinsic (autogenic) process such as bedform migration and
represented by cycles bounded by points of minimum aridity external (allogenic) processes such as climate change and tec-
controlled by regional changes in climate (Fig. 69) has greatly tonic basin evolution, and to demonstrate the extent to which
enhanced the ability to correlate within the basin and has these two sets of processes are independent of each other.
significantly reduced uncertainty in reservoir prediction. Questions not yet fully answered include: What is the preserved
facies response to climatically driven erg-margin expansion and
POTENTIAL FUTURE DEVELOPMENTS contraction? Is it possible to recognize the signature of climate
change in data sets of limited extent or resolution (e.g., core or
Significant advancements in our understanding of the dy- wireline log), and can these signatures be used as a correlation
namic sedimentology and stratigraphy of eolian systems have tool? What is the nature and preserved expression of the com-
been made in recent years. Research throughout the 1990s con- plex interactions that occur between fundamental controlling
centrated on relating the morphology and behavior of modern parameters such as sediment availability, sand saturation of the
eolian systems to the stratigraphic architecture of their counter- airflow, and angle of climb? How do coastal eolian successions
parts preserved in the ancient rock record. Many of the hypotheti- respond to changes in relative sea level, and can we relate
cal models proposed in the late 1970s and early 1980s have now sequence stratigraphic models developed for marine systems to
been updated and refined as new outcrop-based datasets have those developed for eolian systems?
been acquired. The current state of knowledge has progressed to In attempting to answer these questions, detailed relation-
a point where conceptual models for construction, accumulation, ships between large-scale architectural elements and regionally
and preservation of eolian systems can be used as a framework significant bounding surfaces will need to established. This will
with which to develop dynamic facies models that account for 3D require the collection of large 3D datasets from outcrop on a scale
spatial complexity and temporal evolution. Although the eolian hitherto not attempted. The employment of 3D architectural data
facies models developed at the onset of the 21st Century are from outcrop analogue studies is especially important for im-
considerably more sophisticated than their predecessors, there proved subsurface prediction in eolian systems because tradi-
are nevertheless a number of important research questions that tional 1D logs and core have only a limited interpretative scope.
remain to be addressed. In attempting to relate the external form and behavior of modern
Future developments in eolian sedimentology are likely to eolian systems to the potential range of preserved architectures,
focus on the continued development of a set of tools that can be the most likely arena for future research will be the Quaternary
applied to unequivocally differentiate between the products of record, which affords relatively good age constraints using a
74 NIGEL P. MOUNTNEY

Transport
Relative water table = intermediate. Mixed dunes and wet interdunes. Wet eolian system.

T1

Relative water table = low. Maximum dune extent. Dry eolian system.

T2

Relative water table = intermediate. Mixed dunes and wet interdunes. Wet eolian system

T3

Relative water table = high. Minimum due extent. Wet sandsheet with small dunes.
T4

Relative water table = intermediate. Mixed dunes and wet interdunes. Wet eolian system.

T5

Relative water table = low. Maximum dune extent. Dry eolian system.

T6

High

Low

Periods of wet-
interdune
sedimentation
Thickness

Absolute water table


Sediment availability
Relative water table
Accumulation surface
Subsidence

T1 T2 T3 T4 T5 T6
Time

FIG. 62.—Dynamic model for dune and interdune architecture resulting from cyclical temporal variation in controlling parameters
in a mixed wet–dry eolian system. Periodic changes in the position of the water table in relation to the accumulation surface drive
cyclical changes in dune–interdune size. For a fixed rate of dune migration, damp/wet interdune expansion and contraction
occurs during periods of accelerating and decelerating rates of relative water-table rise, respectively.
EOLIAN FACIES MODELS 75

A Transport
Expansion of
damp–wet
interdunes at
expense of dunes
promotes the
preservation of
thicker damp–wet
interdune units.

Black arrows
indicate the former
positions of the
depositional surface
at the time of
B Transport maximum interdune
extent.

Gray arrows
indicate former
positions of the
depositional surface
at the time of
minimum interdune
extent.

FIG. 63.—Models for dune and interdune architecture resulting from cyclical temporal variation in controlling parameters in a wet
eolian system. A) The effects of periodic changes in relative dune–interdune size. For a fixed rate of dune migration, damp and/
or wet interdune expansion and contraction occur during periods of accelerating and decelerating rates of relative water-table
rise, respectively. B) The combined effects of cyclical changes in relative dune–interdune size and related changes in angle of
climb. Angle of climb steepens as the rate of relative water-table rise increases and falls to zero as relative water table becomes
static. Modified after Mountney and Jagger (2004).

variety of recently developed techniques. The continued devel- AHLBRANDT, T.S., ANDREWS, S., AND GWYNNE, D.T., 1978, Bioturbation in
opment of forward numerical stratigraphic models is also likely eolian deposits: Journal of Sedimentary Petrology, v. 48, p. 839–848.
to be a valuable research tool. In particular, the incorporation of ALSHARHAN, A.S., GLENNIE, K., WHITTLE, G.L., AND KENDALL, C.G.ST.C., 1998,
tools for investigating the effects of various configurations of Quaternary Deserts and Climate Change: Rotterdam, Balkema, Pro-
controlling parameters at an erg scale should prove fruitful in ceedings of the International Conference on Quaternary Deserts and
determining the sensitivity of eolian systems to environmental Climate Change, 621 p.
change. ANDERSON, R.S., 1987, A theoretical model for aeolian impact ripples:
Sedimentology, v. 34, p. 943–956.
ACKNOWLEDGMENTS ANDERSON, R.S., 1988, The pattern of grainfall deposition in the lee of
aeolian dunes: Sedimentology, v. 35, p. 175–188.
Alison Jagger is thanked for help in the preparation of some ANDREWS, S., 1981, Sedimentology of Great Sand Dunes, Colorado, in
of the figures used in this paper. John Collinson provided some Ethridge, F.P., and Flores, R.M., eds., Recent and Ancient Non
of the photographs. I am grateful to John Howell and an anony- marine Depositional Environments: Models for Exploration: Soci-
mous reviewer for their encouraging reviews and to Henry ety of Economic Paleontologists and Mineralogists, Special Publica-
Posamentier and Roger Walker for their enthusiasm for the tion 31, p. 279–291.
“Facies Models Revisited” project. BAGNOLD, R.A., 1941, The Physics of Blown Sand and Desert Dunes:
London, Methuen & Company, 265 p.
REFERENCES BATEMAN, M.D., HOLMES, P.J., CARR, A.S., HORTON, B.P., AND JAISWAL, M.K.,
2004, Aeolianite and barrier dune construction spanning the last
AHLBRANDT, T.S., AND FRYBERGER, S.G., 1980, Eolian Deposits in the two glacial–interglacial cycles from the southern Cape coast, South
Nebraska Sand Hills: U.S. Geological Survey, Professional Paper Africa: Quaternary Science Reviews, v. 23, p. 1681–1698.
1120-A, 24 p. BATEMAN, M.D., THOMAS, S.G., AND SINGHVI, A.K., 2003, Extending the
AHLBRANDT, T.S., AND FRYBERGER, S.G., 1981, Sedimentary features and aridity record of the Southwest Kalahari: current problems and
significance of interdune deposits, in Ethridge, F.G., and Flore, R.M., future perspectives: Quaternary International, v. 111, p. 37–49.
eds., Recent and Ancient Non-Marine Depositional Environments: BENAN, C.A.A., AND KOCUREK, G., 2000, Catastrophic flooding of an
Models for Exploration: Society of Economic Paleontologists and aeolian dune field: Jurassic Entrada and Todilto Formations, Ghost
Mineralogists, Special Publication 31, p. 293–314. Ranch, New Mexico, USA: Sedimentology, v. 47, p. 1069–1080.
76 NIGEL P. MOUNTNEY

A Erg Center Transport Erg Margin


Expansion of
damp/–wet
interdunes at
T5 T5
expense of
T4 T4 dunes
T3 T3 promotes the
T2 T2 preservation of
thicker
T1 T1 damp–wet
interdune units.

High High

Low Low
Absolute water table
Periods of wet- Periods of wet-
interdune interdune Sediment availability
sedimentation sedimentation
Relative water table
Thickness

Thickness

Period of 100% Accumulation surface


dune cover
(dry system) Eolian bypass Subsidence
supersurface
generation

B T1 T2 T3
Time
T4 T5
C T1 T2 T3
Time
T4 T5

FIG. 64.—Model for dune and interdune architecture resulting from spatial and temporal variation in controlling parameters. A)
Model incorporating temporal variations of relative dune–interdune size and angle of climb in response to cyclical changes in the
rate of relative water-table change and sediment availability. Additionally, the model incorporates the effects of a spatial decrease
in dune coverage from 100% eolian bedform cover in the erg center to 0% at the outer limit of the erg margin. B, C) Diagrams
illustrating possible configuration of controlling parameters responsible for generating the preserved succession in the erg-center
and erg-margin areas, respectively.

BESLER, H., 1982, The north-eastern Rub’ al Khali within the borders of the Reid I., eds., Desert Sediments; Ancient and Modern: Geological
United Arab Emirates: Zeitschrift für Geomorphologie, v. 26, p. 495–505. Society of London, Special Publication 35, p. 337–359.
B LAKEY , R.C., 1988a, Superscoops—their significance as elements of BRISTOW, C., PUGH, J., AND GOODALL, T., 1996, Internal structure of aeolian
eolian architecture: Geology, v. 16, p. 483–487. dunes in Abu Dhabi determined using ground-penetrating radar:
B LAKEY , R.C., 1988b, Basin tectonics and erg response: Sedimentary Sedimentology, v. 43, p. 995–1003.
Geology, v. 56, p. 127–151. BRISTOW, C.S., BAILEY, S.D., AND LANCASTER, N., 2000, The sedimentary
B LAKEY , R.C., H AVHOLM , K.G., AND JONES , L.S., 1996, Stratigraphic structure of linear sand dunes: Nature, v. 406, p. 56–59.
analysis of eolian interactions with marine and fluvial deposits, BRISTOW, C.S., LANCASTER, N., AND DULLER, G.A.T., 2005, Combining
Middle Jurassic Page Sandstone and Carmel Formation, Colo- ground penetrating radar (GPR) surveys and optical dating to
rado Plateau, USA: Journal of Sedimentary Research, v. 66, p. determine dune migration in Namibia: Geological Society of Lon-
324–342. don, Journal, v. 162, p. 315–321.
B LAKEY , R.C., PETERSON , F., AND KOCUREK , G., 1988, Synthesis of late BROOKFIELD, M.E., 1970, Dune trends and wind regime in central Austra-
Paleozoic and Mesozoic eolian deposits of the western interior of lia: Zeitschrift für Geomorphologie, Supplementband, v. 10, p. 121–
the United States: Sedimentary Geology, v. 56, p. 3–125. 153.
B REED, C.S., AND G ROW, T., 1979, Morphology and distribution of BROOKFIELD, M.E., 1977, The origin of bounding surfaces in ancient
dunes in sand seas observed by remote sensing, in McKee, E.D., aeolian sandstones: Sedimentology, v. 24, p. 303–332.
ed., A Study of Global Sand Seas: U.S. Geological Survey, Profes- CARR-CRABAUGH, M., AND KOCUREK, G., 1998, Continental sequence stratig-
sional Paper 1052, p. 253–302. raphy of a wet eolian system: A key to relative sea-level change, in
B REED, C.S., F RYBERGER, S.G., A NDREWS , S., MC CAULEY , C., L ENNARTZ, F., Shanley, K., and McCabe, P., eds., Relative Roles of Eustasy, Cli-
G EBEL, D., AND H ORSTMAN, K., 1979, Regional studies of sand seas, mate, and Tectonism in Continental Rocks: SEPM, Special Publica-
using Landsat (ERTS) imagery, in McKee, E.D., ed., A Study of tion 59, p. 213–228.
Global Sand Seas: U.S. Geological Survey, Professional Paper CHAKRABORTY, T., AND CHAKRABORTY, C., 2001, Eolian–aqueous interac-
1052, p. 305–397. tions in the development of a Proterozoic sand sheet: Shikaoda
BREED, C.S., MCCAULEY, J.F., AND DAVIS, P.A., 1987, Sand sheets of the Formation, Hosangabad, India: Journal of Sedimentary Research, v.
eastern Sahara and ripple blankets on Mars, in Frostick, L.E., and 71, p. 107–117.
R I R
>121o
R >118o
I I
I
R
R
I I
I
R R R
Facies Association 1 Facies Association 6
2 4
I
Erg-center m Outer-erg- m

27
o
00

0
o
m 0 m 75 0

0
0 50 0
architecture margin architecture
Dynamic facies

1
o
09

80
o
0
model
Mean eolian
bedform
migration Mean fluvial
> 125o transport
direction
F > 225o

F F
D F
Facies Association 2 F Facies Association 5
FD
F

20
m
0 0 20
Facies Association 3 km Facies Association 4
Individual bedforms and sets not drawn to scale
EOLIAN FACIES MODELS

I
Controlling parameters — Inner-erg-margin Controlling parameters — Outer-erg-margin
R
R I R
I
R R I T6
I I T5 T7/8
R T4 T6
R R
T3 T5
I 2 4
T2 T3/4
I I m m
Inner erg T2 Outer-erg-
0 m 50 0 T1 T1 0 m 75 0
margin architecture High High
margin architecture

Low Low
Periods of wet- Periods of wet- Legend
interdune interdune
sedimentation sedimentation Eolian dune Chert Interdune
Absolute water table I surface
Sediment availability Dry interdune Limestone
R
Reactivation
Relative water table Eolian bypass surface

Thickness
Thickness

Sandsheet Paleosol
Accumulation surface & flood surface
generation Flood
Subsidence Damp interdune Fluvial channel F surface
T1 T2 T3 T4 T5 T6 T1 T2 T3 T4 T5 T6 T7 T8 Wet interdune Underlying Deflation
Time Time sequence D surface

FIG. 65.—Integrated dynamic facies model for an erg margin exhibiting dry to wet spatial and temporal variations as a consequence of changes in water table and resultant
changes in sediment availability and angle of climb. Model encapsulates 4D complexity and accounts for facies variability on a range of scales. Based on observations
from the Permian Cedar Mesa Sandstone, southeast Utah, U.S.A.
77
78 NIGEL P. MOUNTNEY

CHRINTZ, T., AND CLEMMENSEN, L.B., 1993, Draa reconstruction, the Permian
Yellow Sands, northeast England, in Pye, K., and Lancaster, N., eds.,
Aeolian Sediments, Ancient and Modern: International Association
of Sedimentologists, Special Publication 16, p. 51–161.
CLEMMENSEN, L.B., 1989, Preservation of interdraa and plinth deposits by
the lateral migration of large linear draas (Lower Permian Yellow
Sands, northeast England): Sedimentary Geology, v. 65, p. 139–151.
CLEMMENSEN, L.B., AND BLAKEY, R.C., 1989, Erg deposits in the Lower
Jurassic Wingate Sandstone, northeastern Arizona—oblique dune
sedimentation: Sedimentology, v. 36, p. 449–470.
CLEMMENSEN, L.B., OLSEN, H., AND BLAKEY, R.C., 1989, Erg-margin depos-
its in the Lower Jurassic Moenave Formation and Wingate Sand-
stone, southern Utah: Geological Society of America, Bulletin, v.
101, p. 759–773.
CLEMMENSEN, L.B., ØXNEVAD, I.E.I., AND DE BOER, P.L., 1994, Climatic
controls on ancient desert sedimentation: some late Palaeozoic
examples from NW Europe and the western interior of the USA, in
de Boer, P.L., and Smith, D.G., eds., Orbital Forcing and Cyclic
Sequences: International Association of Sedimentologists, Special
Publication 19, p. 439–457.
COLLINSON, J.C., 1994, Sedimentary deformational structures, in Maltman,
A., ed., The Geological Deformation of Sediments: London, Chapman
& Hall, p. 95–125.
COOKE, R., WARREN, A., AND GOUDIE, A., 1993, Desert Geomorphology:
London, University College London Press, 526 p.
CORBETT, I., 1993, The modern and ancient pattern of sandflow through the
southwestern Namib deflation basin, in Pye, K., and Lancaster, N.,
eds., Aeolian Sediments, Ancient and Modern: International Associa-
tion of Sedimentologists, Special Publication 16, p. 45–60.
CRABAUGH, M., AND KOCUREK, G., 1993, Entrada Sandstone: An example of
a wet aeolian system, in Pye, K., ed., The Dynamics and Environmen-
tal Context of Aeolian Sedimentary Systems: Geological Society of
London, Special Publication 72, p. 103–126.
DECELLES, P.G., LANGFORD, R.P., AND SCHWARTZ, R.K., 1983, Two new
methods of paleocurrent determination from trough cross-stratifica-
tion: Journal of Sedimentary Petrology, v. 53, p. 629–642.
DOE, T.W., AND DOTT, R.H., JR., 1980, Genetic significance of deformed
cross bedding—with examples from the Navajo and Weber Sand-
stones of Utah: Journal of Sedimentary Petrology, v. 50, p. 793–811.
DOORNCAMP, J.C., BRUNSDEN, D., AND JONES, D.K.C., 1980, Geology, Geomor-
phology and Pedology of Bahrain: Norwich, U.K., Geobooks, 443 p.
DRIESE, S.G., 1985, Interdune pond carbonates, Weber Sandstone (Penn-
5 cm sylvanian–Permian), northern Utah and Colorado: Journal of Sedi-
mentary Petrology, v. 55, p. 187–195.
ELLWOOD, J.M., EVANS, P.D., AND WILSON, I.G., 1975, Small scale aeolian
FIG. 66.—Example of eolian facies in core. Grainflow strata within bedforms: Journal of Sedimentary Petrology, v. 45, p. 554–561.
a cross-stratified eolian dune set overlain by low-angle wind- ESCHNER, T.B., AND KOCUREK, G., 1986, Marine destruction of eolian sand
ripple strata within a dry interdune unit. Leman Sandstone seas—origin of mass flows: Journal of Sedimentary Petrology, v. 56,
Formation, Rotliegend, Permian, southern North Sea. The p. 401–411.
pattern of cross stratification is revealed by the presence of ESCHNER, T.B., AND KOCUREK, G., 1988, Origins of relief along contacts
finer-grained, more tightly packed grainfall deposits between between eolian sandstones and overlying marine strata: American
individual grainflow laminae. Association of Petroleum Geologists, Bulletin, v. 72, p. 932–943.
EVANS, G., KENDALL, C.G.ST.C., AND SKIPWITH, P., 1964, Origin of coastal
flats, the sabkha of the Trucial Coast, Persian Gulf: Nature, v. 202, p.
CHAN, M.A., AND ARCHER, A.W., 1999, Spectral analysis of eolian foreset 759–761.
periodicities—implications for Jurassic decadal-scale paleoclimatic FRANK, A., AND KOCUREK, G., 1996, Toward a model for airflow on the lee
oscillators: Palaeoclimates, v. 3, p. 239–255. side of aeolian dunes: Sedimentology, v. 43, 451–458.
CHAN, M.A., AND ARCHER, A.W., 2000, Cyclic eolian stratification on the FRYBERGER, S.G., 1978, Techniques for the evaluation of surface wind data
Jurassic Navajo Sandstone, Zion National Park: periodicities and in terms of eolian sand drift: U.S. Geological Survey, Open File Report
implications for paleoclimate, in Sprinkel, D.A., Chidsey, T.C., and 78-405, 25 p.
Anderson P.B., eds., Geology of Utah’s Parks and Monuments: Utah FRYBERGER, S.G., 1979, Dune forms and wind regime, in McKee, E.D., ed.,
Geological Association, Publication 28, p. 607–617. A Study of Global Sand Seas: U.S. Geological Survey, Professional
CHAN, M.A., AND KOCUREK, G., 1988, Complexities in eolian and marine Paper 1052, p. 137–169.
interactions—processes and eustatic controls on erg development: FRYBERGER, S.G., 1990a, Great Sand Dunes depositional system—an over-
Sedimentary Geology, v. 56, p. 283–300. view, in Fryberger, S.G., Krystinik, L.F., and Schenk, C.J., eds., Mod-
EOLIAN FACIES MODELS 79

Deposition

Deposition

Deposition
Erg center Basin-margin
Erosion

Erosion

Erosion
Lake–Sabkha and margin system Absolute climate Cycle Facies
Association
Wet Dry
Eolian dune

Interdune

Wetting upward
Deflation surface

Proximal alluvial fan

Medial alluvial fan

Drying upward
Distal alluvial fan

Lake-margin sabkha

Lacustrine

All logs idealized, no vertical scale implied

FIG. 67.—Preserved expression of idealized climatic cycle within three separate sub-environments encountered in the Permian Gas
Basin, southern North Sea. Modified after Howell (1992) and Howell and Mountney (1997).

ern and Ancient Aeolian Deposits: Petroleum Exploration and Pro- FRYBERGER, S.G., SCHENK, C.J., AND KRYSTINIK, L.F., 1988, Stokes surfaces and
duction: SEPM, Denver, p. 1-1–1-9. the effects of near-surface groundwater-table on aeolian deposition:
FRYBERGER, S.G., 1990b, Role of water in eolian deposition, in Fryberger, Sedimentology, v. 35, p. 21–41.
S.G., Krystinik, L.F., and Schenk, C.J., eds., Modern and Ancient GARCÍA-HIDALGO, J.F., TEMIÑO, J., AND SEGURA, M., 2002, Holocene eolian
Eolian Deposits: Petroleum Exploration and Production: SEPM, Den- sediments on the southern border of the Duero Basin (Spain): origin
ver, p. 5-1–5-11. and development of an eolian system in a temperate zone: Journal of
FRYBERGER, S.G., 1990c, Coastal eolian deposits of Oregon, USA, Guererro Sedimentary Research, v. 72, p. 30–39.
Negro, Mexico and Jafurah Sand Sea, Saudi Arabia, in Fryberger, S.G. GEORGE, G.T., AND BERRY, J.K., 1993, A new lithostratigraphy and deposi-
Krystinik, L.F., and Schenk, C.J., eds., Modern and Ancient Eolian tional model for the Upper Rotliegend of the UK sector of the
Deposits: Petroleum Exploration and Production: SEPM, Denver, p. Southern North Sea, in North, C.P., and Prosser, D.J., eds., Character-
11-1–11-15. ization of Fluvial and Aeolian Reservoirs: Geological Society of
FRYBERGER, S.G., 1993, A review of aeolian bounding surfaces, with ex- London, Special Publication 73, p. 291–319.
amples from the Permian Minnelusa Formation, USA, in North, C.P., GEORGE, G.T., AND BERRY, J.K., 1997, Permian (Upper Rotliegend) synsedi-
and Prosser, J.D., eds., Characterization of Fluvial and Aeolian Reser- mentary tectonics, basin development and palaeogeography of the
voirs: Geological Society of London, Special Publication 73, p. 167– southern North Sea, in Ziegler, K., Turner, P., and Daines, S.R., eds.,
197. Petroleum Geology of the Southern North Sea: Future Potential:
FRYBERGER, S.G., AND SCHENK, C., 1981, Wind sedimentation tunnel experi- Geological Society of London, Special Publication 123, p. 31–61.
ments on the origins of aeolian strata: Sedimentology, v. 28, p. 805– GLENNIE, K.W., 1970, Desert Sedimentary Environments: Amsterdam,
821. Elsevier, Developments in Sedimentology, 14, 222 p.
FRYBERGER, S.G., AND SCHENK, C.J., 1988, Pin stripe lamination—a distinc- GLENNIE, K.W., 1972, Permian Rotliegendes of North West Europe
tive feature of modern and ancient eolian sediments: Sedimentary interpreted in light of modern desert sedimentation studies: Ameri-
Geology, v. 55, p. 1–15. can Association of Petroleum Geologists, Bulletin, v. 56, p. 1048–
FRYBERGER, S.G., AL-SARI, A.M., CLISHAM, T.J., RIZVI, S.A.R., AND AL-HINAI, 1071.
K.G., 1984, Wind sedimentation in the Jafurah sand sea, Saudi Arabia: GLENNIE, K.W., 1990, Lower Permian—Rotliegend, in Glennie, K.W., ed.,
Sedimentology, v. 31, p. 413–431. Introduction to the Petroleum Geology of the North Sea: Oxford,
FRYBERGER, S.G., AHLBRANDT, T.S., AND ANDREWS, S., 1979, Origin, sedimen- U.K., Blackwell Science, p. 120–152.
tary features, and significance of low-angle eolian “sand sheet” GLENNIE, K.W., AND BULLER, A.T., 1983, The Permian Weissliegend of NW
deposits, Great Sand Dunes National Monument and vicinity, Colo- Europe—the partial deformation of aeolian dune sands caused by the
rado: Journal of Sedimentary Petrology, v. 49, p. 733–746. Zechstein Transgression: Sedimentary Geology, v. 35, p. 43–81.
FRYBERGER, S.G., AL-SARI, A.M., AND CLISHAM, T.J., 1983, Eolian dune, GREELEY, R., AND IVERSON, J.D., 1985, Wind as a Geological Process on Earth,
interdune, sand sheet, and siciliclastic sabkha sediments of an off- Mars, Venus and Titan: Cambridge, U.K., Cambridge University
shore prograding sand sea, Dharan Area, Saudi Arabia: American Press, 345 p.
Association of Petroleum Geologists, Bulletin, v. 67, p. 280–312. GUNATILAKA, A., AND MWANGO, S., 1987, Continental sabkha pans and
FRYBERGER, S.G., HESP, P., AND HASTINGS, K., 1992, Aeolian granule ripple associated nebkhas in southern Kuwait, Arabian Gulf, in Frostick,
deposits, Namibia: Sedimentology, v. 39, p. 319–331. L.E., and Reid, I., eds., Desert Sediments, Ancient and Modern:
FRYBERGER, S.G., KRYSTINIK, L.F., AND SCHENK, C.J, 1990, Tidally flooded Geological Society of London, Special Publication 35, p. 187–203.
back-barrier dunefield, Guerrero Negro area, Baja California, Mexico: HASIOTIS, S.T., 2002, Continental Trace Fossils: SEPM, Short Course Notes,
Sedimentology, v. 37, p. 23–43. no. 51, 132 p.
80 NIGEL P. MOUNTNEY

HERRIES, R.D., 1992, Sedimentology of Continental Erg Margin Interac-


A Eolian erg and erg margin Lithology Preserved
climatic cycle
tions: Unpublished Ph.D. Thesis, University of Aberdeen, 205 p.
HERRIES, R.D., 1993, Contrasting styles of fluvial–aeolian interaction at a
Subsidence
DEPTH/THICKNESS

downwind erg margin: Jurassic Kayenta–Navajo transition, north-


eastern Arizona, USA, in North, C.P., and Prosser, J.D., eds., Charac-
Accumulation
Drying- terization of Fluvial and Aeolian Reservoirs: Geological Society of
upward
Supersurface
cycles
London, Special Publication 73, p. 199–218.
HERRIES, R.D., AND COWAN, G., 1997, Challenging the ‘sheetflood’ myth: the
role of water-table-controlled sabkha deposits in redefining the depo-
sitional model for the Ormskirk Sandstone Formation (Lower Trias-
Sediment availability
sic), East Irish Sea Basin, in Meadows, N.S., Trueblood, S.P., Hardman,
RATE

Subsidence M., and Cowan, G., Petroleum Geology of the Irish Sea and Adjacent
eolian dune slipface deposits Areas: Geological Society of London, Special Publication 124, p. 253–
Wet Dry Wet Dry Wet 276.
wind-rippled sandsheet deposits
TIME
HOROWITZ, D.H., 1982, Geometry and origin of large-scale deformation
Climate structures in some ancient wind-blown sand deposits: Sedimentol-
ogy, v. 29, p. 155–180.
Preserved HOWELL, J.A., 1992, Sedimentology of the Rotliegend Supergroup of the
B Basin-margin alluvial fan Lithology
climatic cycle
UK Southern North Sea: Unpublished Ph.D. Thesis, University of
DEPTH/THICKNESS

Wetting- Birmingham, 345 p.


upward
cycle
HOWELL, J., AND MOUNTNEY, N., 2001, Aeolian grain flow architecture: hard
data for reservoir models and implications for red bed sequence
Drying– stratigraphy: Petroleum Geoscience, v. 7, p. 51–56.
wetting
Cycle
HOWELL, J.A., AND MOUNTNEY, N.P., 1997, Climatic cyclicity and accommo-
dation space in arid to semi-arid depositional systems: An example
Proximal alluvial-fan deposits from the Rotliegend Group of the Southern North Sea, in Ziegler, K.,
RATE

Turner, P., and Daines, S.R., eds., Petroleum Geology of the Southern
Intrabasinal fluvial deposits
North Sea: Future Potential: Geological Society of London, Special
Wet Dry Wet Dry Wet
Deflation lag deposits Publication 123, p. 63–86.
TIME HUMMEL, G., AND KOCUREK, G., 1984, Interdune areas of the back-Island
Climate dune field, North Padre Island, Texas: Sedimentary Geology, v. 39, p.
1–26.
HUNTER, R.E., 1977, Basic types of stratification in small eolian dunes:
C Lake and lake-margin sabkha Lithology Preserved
climatic cycle
Sedimentology, v. 24, p. 361–387.
HUNTER, R.E., 1981, Stratification styles in eolian sandstones: Some Penn-
DEPTH/THICKNESS

Drying- sylvanian to Jurassic examples from the western interior USA, in


upward
cycle Ethridge, F.G., and Flores, R.M., eds., Recent and Ancient Non-
Marine Depositional Environments, Models for Exploration: Society
Symmetrical
of Economic Paleontologists and Mineralogists, Special Publication
drying–
wetting 31, p. 315–329.
cycle HUNTER, R.E., 1985, A kinematic model for the structure of lee-side
deposits: Sedimentology, v. 32, p. 409–422.
RATE

Playa-lake mudstone

Lake-margin sabkha deposits


HUNTER, R.E., AND RICHMOND, B.M., 1988, Daily cycles in coastal dunes:
Wind-rippled eolian sandsheet
Sedimentary Geology, v. 55, p. 43–67.
Wet Dry Wet Dry Wet HUNTER, R.E., AND RUBIN, D.M., 1983, Interpreting cyclic cross-bedding,
TIME
with an example from the Navajo Sandstone, in Brookfield, M.E., and
Climate Ahlbrandt, T.S., eds., Eolian Sediments and Processes: Amsterdam,
Elsevier, Developments in Sedimentology, no. 38, p. 429–454.
FIG. 68.—Diagrams illustrating how variations in climate control HUNTER, R.E., RICHMOND, B.M., AND ALPHA, T.R., 1983, Storm-controlled
the preserved expression in the Permian Rotliegend Group, oblique dunes of the Oregon Coast: Geological Society of America,
Southern North Sea. A) Eolian erg and erg margin. B) Basin- Bulletin, v. 94, p. 1450–1465.
margin alluvial fan. C) Lake and lake-margin sabkha. Modi- HUNTINGTON, E., 1907, Some characteristics of the glacial period in non-
fied after Howell and Mountney (1997). glaciated regions: Geological Society of America, Bulletin, v. 18, p.
351–388.
HUNTOON, J.E., AND CHAN, M.A., 1987, Marine origin of paleotopographic
relief on eolian White Rim Sandstone (Permian), Elaterite Basin, Utah:
HAVHOLM, K.G., AND KOCUREK, G., 1994, Factors controlling aeolian se- American Association of Petroleum Geologists, Bulletin, v. 71, p.
quence stratigraphy: clues from super bounding surface in the Middle 1035–1045.
Jurassic Page Sandstone: Sedimentology, v. 41, p. 913–934. JAGGER, A., 2003, Sedimentology and Stratigraphic Evolution of the Per-
HAVHOLM, K.G., BLAKEY, R.C., CAPPS, M., JONES, L.S., KING, D.D., AND mian Cedar Mesa Sandstone, SE Utah: Unpublished Ph.D. thesis,
KOCUREK, G., 1993, Aeolian genetic stratigraphy: an example from University of Keele, 391 p.
the Middle Jurassic Page Sandstone, Colorado Plateau, in Pye, K., JERRAM, D.A., MOUNTNEY, N.P., AND STOLLHOFEN, H., 1999, Facies architec-
and Lancaster, N., eds., Aeolian Sediments, Ancient and Modern: ture of the Etjo Sandstone Formation and its interaction with the
International Association of Sedimentologists, Special Publication basal Etendeka flood basalts of NW Namibia: Implications for
16, p. 87–107. offshore analogues, in Cameron, N., Bate, R., and Clure, V., eds., Oil
EOLIAN FACIES MODELS 81

Northern Raven- North East


Audrey Indefatigable Leman Hewett Anglia Sole Pit spurn Midland Shelf Amethyst

12

11 Lithostratigraphy
Upper Leman Sand Fm.
10
Silverpit Fm.
9 Leman Sand Fm.
8
7 Relative Climate
WET - DRY

6 Unit Number
8
Unit boundary
5 (Chronostratigraphic
7 surface)
4
Each curve is a summary of the relative
3 climatic cycles from several study wells.
Note the varied expression of single
cycles and that the thickness of each
2 cycle is typically constant within
individual structural blocks. The unit
cycle boundaries are time lines and
1 cross-cut the lithostratigraphic formation
boundaries.

FIG. 69.—Correlation of climatic cycles across the Permian Rotliegend Group, southern North Sea. Modified after Howell (1992) and
Howell and Mountney (1997). Each curve is a summary of the relative climatic cycles from several study wells. Note the varied
expression of single cycles and that the thickness of each cycle is typically constant within individual structural blocks. The unit
cycle boundaries are time lines and crosscut the lithostratigraphic formation boundaries.

and Gas Habitats of the South Atlantic: Geological Society of Lon- and Stokes, S., eds., Aeolian Environments, Sediments, and Land-
don, Special Publication 153, p. 367–380. forms: Chichester, U.K., John Wiley & Sons Ltd., p. 239–259.
JERRAM , D.A., MOUNTNEY , N., H OWELL, J., AND S TOLLHOFEN, H., 2000, The KOCUREK, G., AND DOTT, R.H., JR., 1981, Distinctions and uses of strati-
fossilized desert: recent developments in our understanding of fication types in the interpretation of eolian sand: Journal of
the Lower Cretaceous deposits in the Huab Basin, NW Namibia: Sedimentary Petrology, v. 51, p. 579–595.
Geological Survey of Namibia, Communications, v. 12, p. 269– KOCUREK, G., AND FIELDER, G., 1982, Adhesion structures: Journal of
278. Sedimentary Petrology, v. 52, p. 1229–1241.
KAMOLA, D.L., AND HUNTOON, J.E., 1994, Changes in rate of transgres- KOCUREK, G., AND HAVHOLM, K.G., 1993, Eolian sequence stratigraphy—
sion across the Permian White Rim Sandstone, southern Utah: a conceptual framework, in Weimer, P., and Posamentier, H.W.,
Journal of Sedimentary Research, v. B64, p. 202–210. eds., Siciliclastic Sequence Stratigraphy: American Association of
KHALAF, F., 1989, Textural characteristics and genesis of the aeolian Petroleum Geologists, Memoir 58, p. 393–409.
sediments in the Kuwaiti Desert: Sedimentology, v. 36, p. 253–271. KOCUREK, G., AND HUNTER, R.E., 1986, Origin of polygonal fractures in
KHALAF, F.I., GHARIB, I.M., AND AL HASHASH, M.Z., 1984, Types and sand, uppermost Navajo and Page Sandstones, Page, Arizona:
characteristics of recent surface deposits of Kuwait, Arabian Gulf: Journal of Sedimentary Petrology, v. 56, p. 895–904.
Journal of Arid Environments, v. 7, p. 9–33. KOCUREK, G., AND NIELSON, J., 1986, Conditions favourable for the
KOCUREK, G., 1981, Significance of interdune deposits and bounding formation of warm-climate aeolian sand sheets: Sedimentology, v.
surfaces in aeolian dune sands: Sedimentology, v. 28, p. 753–780. 33, p. 795–816.
KOCUREK, G., 1984, Origin of first-order bounding surfaces in aeolian KOCUREK, G., AND LANCASTER, N., 1999, Aeolian system sediment state:
sandstones—Reply: Sedimentology, v. 31, p. 125–127. theory and Mojave Desert Kelso dune field example: Sedimentol-
KOCUREK, G., 1988, First-order and super bounding surfaces in eolian ogy, v. 46, p. 505–515.
sequences—bounding surfaces revisited: Sedimentary Geology, v. KOCUREK, G., HAVHOLM, K.G., DEYNOUX, M., AND BLAKEY, R.C., 1991a,
56, p. 193–206. Amalgamated accumulations resulting from climatic and eustatic
KOCUREK, G., 1991, Interpretation of ancient eolian sand dunes: Annual changes, Akchar Erg, Mauritania: Sedimentology, v. 38, p. 751–772.
Review of Earth and Planetary Sciences, v. 19, p. 43–75. KOCUREK, G., KNIGHT, J., AND HAVHOLM, K., 1991b, Outcrop and semi-regional
KOCUREK, G., 1996, Desert aeolian systems, in Reading, H.G., ed., Sedi- three-dimensional architecture and reconstruction of a portion of the
mentary Environments; Processes, Facies and Stratigraphy, Third eolian Page Sandstone (Jurassic), in Miall, A.D., and Tyler, N., eds., The
Edition: Oxford, U.K., Blackwell Science, p. 125–153. Three-Dimensional Facies Architecture of Terrigenous Clastic Sedi-
KOCUREK, G., 1999, The aeolian rock record (Yes, Virginia, it exists but it ments and Its Implications for Hydrocarbon Discovery and Recovery:
really is rather special to create one), in Goudie, A.S., Livingstone, I., SEPM, Concepts in Sedimentology and Paleontology, v. 3, p. 25–43.
82 NIGEL P. MOUNTNEY

KOCUREK, G., TOWNSLEY, M., YEH, E., HAVHOLM, K., AND SWEET, M.L., 1992, MCKEE, E.D., AND MOIOLA, R.J., 1975, Geometry and growth of the White
Dune and dune-field development on Padre Island, Texas, with Sands dune field, New Mexico: U.S. Geological Survey, Journal of
implications for interdune deposition and water-table-controlled Research, v. 3, p. 59–66.
accumulation: Journal of Sedimentary Petrology, v. 62, p. 622–635. MCKEE, E.D., DOUGLASS, J.R., AND RITTENHOUSE, S., 1971, Deformation of lee
KOCUREK, G., LANCASTER, N., CARR, M., AND FRANK, A., 1999, Tertiary side laminae in eolian dunes: Geological Society of America, Bulle-
Tsondab Sandstone Formation: preliminary bedform reconstruction tin, v. 82, p. 359–378.
and comparison to modern Namib Sand Sea dunes: Journal of African MIDDLETON, L.T., AND BLAKEY, R.C., 1983, Processes and controls on the
Earth Sciences, v. 29, p. 629–642. intertonguing of the Kayenta and Navajo formations, northern
KOCUREK, G., ROBINSON, N.I., AND SHARP, J.M., JR., 2001, The response of the Arizona: Eolian–fluvial interactions, in Brookfield, M.E., and
water table in coastal aeolian systems to changes in sea level: Sedi- Ahlbrandt, T.S., eds., Eolian Sediments and Processes: Amsterdam,
mentary Geology, v. 139, p. 1–13. Elsevier, Developments in Sedimentology, no. 38, p. 613–634.
KOSTER, E.A., 1988, Ancient and modern cold-climate aeolian sand depo- MOUNTNEY, N., AND HOWELL, J., 2000, Aeolian architecture, bedform
sition: a review: Journal of Quaternary Science, v. 3, p. 69–83. climbing and preservation space in the Cretaceous Etjo Formation,
LANCASTER, N., 1985, Winds and sand movements in the Namib sand sea: NW Namibia: Sedimentology, v. 47, p. 825–849.
Earth Surface Processes and Landforms, v. 10, p. 607–619. MOUNTNEY, N.P., AND JAGGER, A., 2004, Stratigraphic evolution of a wet
LANCASTER, N., 1986, Grain-size characteristics of linear dunes in the south- aeolian system: The Permian Cedar Mesa Sandstone, SE Utah:
west Kalahari: Journal of Sedimentary Petrology, v. 56, p. 395–400. Sedimentology, v. 51, p. 713–743.
LANCASTER, N., 1988, Controls on eolian dune size and spacing: Geology, MOUNTNEY, N.P., AND RUSSELL, A.J., 2004, Sedimentology of aeolian
v. 16, p. 972–975. sandsheet deposits in the Askja region of northeast Iceland: Sedi-
LANCASTER, N., 1995, Geomorphology of Desert Dunes: New York, mentary Geology, v. 166, p. 223–244.
Routledge, 290 p. MOUNTNEY, N.P., AND THOMPSON, D.B., 2002, Stratigraphic evolution and
LANCASTER, N., 1998, Dune morphology, chronology and Quaternary preservation of aeolian dune and damp/wet interdune strata: an
climate change, in Alsharhan, A.S., Glennie, K., Whittle, G.L., and example from the Triassic Helsby Sandstone Formation, Cheshire
Kendall, C.G.St.C., eds., Quaternary Deserts and Climate Change: Basin, UK: Sedimentology, v. 49, p. 805–834.
Proceedings of the International Conference on Quaternary Deserts MOUNTNEY, N.P., HOWELL, J.A., FLINT, S.S., AND JERRAM, D.A., 1998, Aeolian
and Climate Change: Rotterdam, Balkema, p. 339–349. and alluvial deposition within the Mesozoic Etjo Sandstone Forma-
LANCASTER, N., AND TELLER, J.T., 1988, Interdune deposits of the Namib tion, North West Namibia: Journal of African Earth Sciences, v. 27,
Sand Sea: Sedimentary Geology, v. 55, p. 91–107. p. 175–192.
LANCASTER, N., GREELEY, R., AND CHRISTENSEN, P.R., 1987, Dunes of the Gran MOUNTNEY, N.P., HOWELL, J.A., FLINT, S.S., AND JERRAM, D.A., 1999a,
Desierto sand sea, Sonora, Mexico: Earth Surface Processes and Climate, sediment supply and tectonics as controls on the deposi-
Landforms, v. 12, p. 277–288. tion and preservation of the aeolian–fluvial Etjo Sandstone Forma-
LANGFORD, R.P., 1989, Fluvial–aeolian interactions: part 1, modern sys- tion, Namibia: Geological Society of London, Journal, v. 156, p. 771–
tems: Sedimentology, v. 36, p. 1023–1035. 779.
LANGFORD, R., AND CHAN, M.A., 1988, Flood surfaces and deflation surfaces MOUNTNEY, N.P., HOWELL, J.A., FLINT, S.S., AND JERRAM, D.A., 1999b,
within the Cutler Formation and Cedar Mesa Sandstone (Permian), Relating eolian bounding-surface geometries to the bed forms that
southeastern Utah: Geological Society of America, Bulletin, v. 100, p. generated them: Etjo Formation, Cretaceous, Namibia: Geology, v.
1541–1549. 27, p. 159–162.
LANGFORD, R.P., AND CHAN, M.A., 1989, Fluvial–aeolian interactions: part NICKLING, W.G., MCKENNA NEUMAN, C., AND LANCASTER, N., 2002, Grainfall
2, ancient systems: Sedimentology, v. 36, p. 1037–1051. processes in the lee of transverse dunes, Silver Peak, Nevada:
LOOPE, D.B., 1984, Eolian origin of Upper Paleozoic sandstones, southeast- Sedimentology, v. 49, p. 191–209.
ern Utah: Journal of Sedimentary Petrology, v. 54, p. 563–580. NIELSON, J., AND KOCUREK, G., 1986, Climbing zibars of the Algodones:
LOOPE, D.B., 1985, Episodic deposition and preservation of eolian sands— Sedimentary Geology, v. 48, p. 1–15.
a late Paleozoic example from southeastern Utah: Geology, v. 13, p. OLSEN, H., DUE, P.H., AND CLEMMENSEN, L.B., 1989, Morphology and
73–76. genesis of asymmetric adhesion warts—a new adhesion surface
LOOPE, D.B., 1988, Rhizoliths and ancient aeolianites: Sedimentary Geol- structure: Sedimentary Geology, v. 61, p. 277–285.
ogy, v. 56, p. 301–314. ØXNEVAD, I.E.I., 1991, Aeolian and mixed aeolian–subaqueous sedimen-
LOOPE, D.B., ROWE, C.M., AND JOECKEL, R.M., 2001, Annual monsoon rains tation in modern and ancient sub-tropical desert basins: examples
recorded by Jurassic dunes: Nature, v. 412, p. 64–66. from the Sahara and Permo-Triassic of NW Europe: Unpublished
LOOPE, D.B., SWINEHART, J.B., AND MASON, J.P., 1995, Dune-dammed Doctor of Science thesis, University of Bergen, Norway, 161 p.
paleovalleys of the Nebraska Sand Hills—intrinsic versus climatic (volume I), 111 p. (volume II).
controls on the accumulation of lake and marsh sediments: Geologi- PORTER, M.L., 1986, Sedimentary record of erg migration: Geology, v. 14,
cal Society of America, Bulletin, v. 107, p. 396–406. p. 497–500.
MAINGUET, M., AND CHEMIN, M.-C., 1983, Sand seas of the Sahara and Sahel: PULVERTAFT, T.C.R., 1985, Aeolian dune and wet interdune sedimenta-
An explanation of their thickness and sand dune type by the sand tion in the Middle Proterozoic Dala Sandstone, Sweden: Sedimen-
budget principle, in Brookfield, M.E., and Ahlbrandt, T.S., eds., tary Geology, v. 44, p. 93–111.
Eolian Sediments and Processes: Amsterdam, Elsevier, Develop- PURSER, B.H., AND EVANS, G., 1973, Regional sedimentation along the
ments in Sedimentology, no. 38, p. 353–363. Trucial Coast, Persian Gulf, in Purser, B.H., ed., The Persian Gulf:
MAXWELL, T.A., AND HAYNES, C.V., 1989, Large-scale, low-amplitude Berlin, Springer, p. 211–231.
bedforms (chevrons) in the Selima Sand Sheet, Egypt: Science, v. 243, PYE, K., 1982, Negatively skewed aeolian sands from a humid tropical
p. 1179–1182. coastal dunefield, northern Australia: Sedimentary Geology, v. 31,
MCKEE, E.D., 1966, Structures of dunes at White Sands National Monu- p. 249–266.
ment, New Mexico (and a comparison with structures of dunes from PYE, K., 1983, Coastal dunes: Progress in Physical Geography, v. 7, p.
other selected areas): Sedimentology, v. 7, p. 1–69. 531–546.
MCKEE, E.D., 1979, A Study of Global Sand Seas: U.S. Geological Survey, PYE, K., AND TSOAR, H., 1987, The mechanics and geological implications of
Professional Paper 1052, 421 p. dust transport and deposition in deserts, with particular reference to
EOLIAN FACIES MODELS 83

loess formation and dune sand diagenesis in the northern Negev, a high-resolution record from the Chott Rharsa basin, Tunisia: The
Israel, in Frostick, L.E., and Reid, I., eds., Desert Sediments; Ancient Holocene, v. 9, p. 141–147.
and Modern: Geological Society of London, Special Publication 35, p. TALBOT, M.R., 1985, Major bounding surfaces in aeolian sandstones—a
139–156. climatic model: Sedimentology, v. 32, p. 257–265.
RANKEY, E.C., 1997, Relations between relative changes in sea level and THOMPSON, D.B., 1969, Dome shaped aeolian dunes in the Frodsham
climate shifts: Pennsylvanian–Permian mixed carbonate–siliciclastic Member of the so-called ‘Keuper’ Sandstone Formation [Scythian–
strata, western United States: Geological Society of America, Bulletin, ?Anisan: Triassic] at Frodsham, Cheshire (England): Sedimentary
v. 109, p. 1089–1100. Geology, v. 3, p. 263–289.
RUBIN, D.M., 1987a, Cross-Bedding, Bedforms and Paleocurrents: SEPM, TREWIN, N.H., 1993, Controls on fluvial deposition in mixed fluvial and
Concepts in Sedimentology and Paleontology, v. 1, 187 p. aeolian facies within the Tumblagooda Sandstone (Late Silurian) of
RUBIN, D.M., 1987b, Formation of scalloped cross-bedding without un- Western Australia: Sedimentary Geology, v. 85, p. 387–400.
steady flows: Journal of Sedimentary Petrology, v. 57, p. 39–45. WARREN, A., 1971, The dunes of the Ténéré Desert: Geographical Journal,
RUBIN, D.M., AND HUNTER, R.E., 1982, Bedform climbing in theory and v. 137, 458–461.
nature: Sedimentology, v. 29, p. 121–138. WASSON, R.J., AND HYDE, R., 1983, Factors determining desert dune type:
RUBIN, D.M., AND HUNTER, R.E., 1983, Reconstructing bedform assem- Nature, v. 304, p. 337–339.
blages from compound crossbedding, in Brookfield, M.E., and WERNER, B.T., AND KOCUREK, G., 1999, Bedform spacing from defect dy-
Ahlbrandt, T.S., eds., Eolian Sediments and Processes: Amsterdam, namics: Geology, v. 27, p. 727–730.
Elsevier, Developments in Sedimentology, no. 38, p. 407–427. WILSON, I.G., 1971, Desert sandflow basins and a model for the develop-
RUBIN, D.M., AND HUNTER, R.E., 1984, Origin of 1st-order bounding surfaces ment of ergs: Geographical Journal, v. 137, p. 180–199.
in aeolian sandstones—Reply: Sedimentology, v. 31, p. 128–132. WILSON, I.G., 1972. Aeolian bedforms—their development and origins:
RUBIN, D.M., AND HUNTER, R.E., 1985, Why deposits of longitudinal dunes Sedimentology, v. 19, p. 173–210.
are rarely recognized in the geologic record: Sedimentology, v. 32, p. WILSON, I.G., 1973, Ergs: Sedimentary Geology, v. 10, p. 77–106.
147–157.
SCHENK, C.J., 1990, Eolian deposits of North Padre Island, Texas, in
Fryberger, S.G., Krystinik, L.F., and Schenk, C.J., eds., Modern and
Ancient Eolian Deposits: Petroleum Exploration and Production:
Denver, SEPM, p. 10-1–10-7.
SCHERER, C.M.S., 2000, Eolian dunes of the Botucatu Formation (Creta-
ceous) in southernmost Brazil: morphology and origin: Sedimentary
Geology, v. 137, p. 63–84.
SCHERER, C.M.S., 2002, Preservation of aeolian genetic units by lava flows
in the Lower Cretaceous of the Paraná Basin, southern Brazil: Sedi-
mentology, v. 49, p. 97–116.
SHARP, R.P., 1963, Wind ripples: Journal of Geology, v. 71, p. 617–636.
SHARP, R.P., 1966, Kelso dunes, Mojave Desert, California: Geological
Society of America, Bulletin, v. 77, p. 1045–1074.
SIMPSON, E.L., AND ERIKSSON, K.A., 1993, Thin eolianites interbedded within
a fluvial and marine succession—Early Proterozoic Whitworth For-
mation, Mount Isa inlier, Australia: Sedimentary Geology, v. 87, p.
39–62.
SIMPSON, E.L., AND LOOPE, D.B., 1985, Amalgamated interdune deposits,
White Sands, New Mexico: Journal of Sedimentary Petrology, v. 55, p.
361–365.
STANISTREET, I.G., AND STOLLHOFEN, H., 2002, Hoanib River flood deposits of
Namib Desert interdunes as analogues for thin permeability barrier
mudstone layers in aeolianite reservoirs: Sedimentology, v. 49, p.
719–736.
STOKES, S., THOMAS, S.G., AND WASHINGTON, R., 1997, Multiple episodes of
aridity in southern Africa since the last interglacial period: Nature, v.
388, p. 154–158.
STOKES, W.L., 1968, Multiple parallel-truncation bedding planes—a fea-
ture of wind deposited sandstone formations: Journal of Sedimentary
Petrology, v. 38, p. 510–515.
SWEET, M.L., 1992, Lee-face air-flow, surface processes, and stratification
types—their significance for refining the use of eolian cross-strata as
paleocurrent indicators: Geological Society of America, Bulletin, v.
104, p. 1528–1538.
SWEET, M.L., AND KOCUREK, G., 1990, An empirical model of aeolian dune
lee-face airflow: Sedimentology, v. 37, p. 1023–1038.
SWEZEY, C., 2001, Eolian sediment responses to late Quaternary climate
changes: temporal and spatial patterns in the Sahara: Palaeogeography
Palaeoclimatology Palaeoecology, v. 167, p. 119–155.
SWEZEY, C., LANCASTER, N., KOCUREK, G., DEYNOUX, M., BLUM, M., PRICE, D.,
AND PION, J.-C., 1999, Response of aeolian systems to Holocene cli-
matic and hydrologic changes on the northern margin of the Sahara:
84 NIGEL P. MOUNTNEY
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 85

FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS

JOHN S. BRIDGE
Department of Geological Sciences, Binghamton University, Binghamton, New York 13902-6000, U.S.A.
e-mail: jbridge@binghamton.edu

ABSTRACT: Recent development of fluvial facies models has been due to improved description of natural river and floodplain processes
and deposits using: (1) ground-penetrating radar (GPR) combined with cores and trenches to describe modern deposits in 3D; (2) study
of frozen rivers to allow easy access to the entire channel belt and procurement of undisturbed cores; (3) optically stimulated luminescence
(OSL) for improved dating of deposits; (4) high-resolution remote sensing over large areas and at short time intervals in order to determine
temporal changes in channel and floodplain geometry due to erosion and deposition; (5) new measuring equipment such as acoustic
Doppler current profilers (ADCP), high-resolution multibeam sonar, and GPS, for measuring surface topography, flow, and sedimentary
processes. However, there is still a lack of studies of river geometry, flow, and sedimentary processes at the all-important high flow stages,
especially on big rivers and floodplains.
Laboratory studies of bed geometry, flow, and sediment transport, erosion, and deposition have been undertaken for a range of scales,
from small bedforms such as ripples, dunes, and antidunes, to bars and channels, to whole channel belt–floodplain systems. Controls on
river and floodplain mechanics such as sediment supply, base level ,and tectonism have also been evaluated. However, there are scaling
problems with laboratory experiments that become more acute as the scale of the system increases.
The new field and laboratory data have allowed development of new qualitative and quantitative fluvial depositional models. Such
models account for the fact that: (1) there are different superimposed scales of fluvial forms and associated stratasets in rivers and
floodplains; (2) the geometry and mode of migration of any scale of fluvial form (e.g., dune, bar, channel, channel-belt) is closely related
to the geometry and internal character of the associated strataset, which allows development of generalized depositional models for the
different scales; (3) changes in flow stage over various time scales affect the nature of deposits. These new models use consistent descriptive
terminology and dispel many of the extant misconceptions about fluvial deposits.
Quantitative, process-based models of fluvial deposits exist, but are not well developed, especially for the longer-term and larger-scale
processes and deposits. Process-based models of the effects of tectonism, climate, and base-level change on fluvial deposits are in their
infancy. Furthermore, most models are difficult to test. These problems with quantitative models are due to lack of appropriate quantitative
data, and difficulties in mathematical modeling of complex natural systems. As a result of this, stochastic models are commonly used to
represent fluvial stratigraphy, given initial data from wells, cores, and geophysical surveys. Development of quantitative models is
essential if we are to understand and predict the nature and spatial distribution of ancient fluvial deposits, and to characterize aquifers and
hydrocarbon reservoirs for subsurface fluid flow simulations. Such development will require more studies of rivers and floodplains during
floods, and more mathematical sophistication.

INTRODUCTION modern alluvium is generally not known well enough. This is


due partly to difficulties in describing deposits below the water
Rivers and floodplains (including alluvial fans and deltas) table, and in studying depositional processes during the all-
are important features of the Earth’s surface, both now and in important high-flow stages and over large time and space scales
the past. Ancient fluvial deposits are indicators of past Earth (see reviews by Bridge, 1985, 1993, 2003). Also, field and labora-
surface environments, and may contain economically impor- tory studies are expensive, and theoretical modeling of complex
tant resources such as water, oil, gas, coal, and placer minerals. fluvial processes is difficult.
Understanding of fluvial sedimentary forms, processes, and Over the past decade or so, difficulties in describing modern
deposits has come from: (1) field studies of modern environ- fluvial deposits have been overcome by: (1) use of ground-
ments; (2) laboratory flume studies using physical models; and penetrating radar (GPR) in combination with cores and trenches
(3) construction of models based on these studies. Depositional to describe different scales of deposits in detail (e.g., Jol and
(facies) models and direct modern analogs allow rational inter- Smith, 1991; Gawthorpe et al., 1993; Huggenberger, 1993; Jol,
pretation of ancient deposits, and can aid prediction of the 1995; Bridge et al., 1995; Bridge et al., 1998; Beres et al., 1995;
nature of subsurface deposits where data (e.g., cores, well logs, Beres et al., 1999; McMechan et al., 1997; Bristow et al., 1999;
seismic) are sparse. Depositional models can be qualitative Szerbiak et al., 2001; Corbeanu et al., 2001; Skelly et al., 2003;
(graphic) and/or quantitative (numerical), static and/or dy- Woodward et al., 2003; Best et al., 2003; Lunt et al., 2004a, 2004b);
namic (forward), stochastic and/or deterministic. Ideally, a (2) study of channel deposits in frozen rivers, allowing easy
depositional model must represent landforms and sedimentary access to the whole channel belt, and the procurement of undis-
processes accurately, must contain detailed sedimentary infor- turbed cores of unconsolidated gravel (Lunt et al., 2004a, 2004b);
mation (including the various superimposed scales of strata), and (3) improved methods of determining the age of fluvial
should be quantitative, and should have some predictive value. deposits, particularly optically stimulated luminescence (Duller,
A depositional model should also provide parameters (e.g., 1996; Aitken, 1998). Also, our ability to document changes in the
permeability, porosity) relevant to modeling fluid flow through geometry of channels and floodplains arising from erosion and
aquifers and hydrocarbon reservoirs. Most fluvial depositional deposition has improved by using aerial photos and satellite
models (e.g., Miall, 1992, 1996; Bridge, 1993, 2003; Collinson, images taken at short time intervals (e.g., Lane et al., 1994; Lane
1996) do not meet these ideals, because the nature and origin of et al., 1995; Lane et al., 1998; Lane et al., 2001; Lane et al., 2003;

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 85–170.
86 JOHN S. BRIDGE

Ashmore and Church, 1998; Stojic et al., 1998; Chandler, 1999; from modern environments. Useful quantitative models are more
Westaway et al., 2000; Ashworth et al., 2000; Best et al., 2003; difficult to construct and test as the time scale and spatial scale
Lunt and Bridge, 2004). These studies have allowed the con- increase. There is a lack of linkage between models treating
struction of a new generation of fluvial depositional models different scales, and there are inadequate data to test long-term,
(e.g., Best et al., 2003; Lunt et al., 2004a, 2004b; Bridge and Lunt, large-scale models. As a result, quantitative fluvial models are in
2006). their infancy (Bridge, in press).
There have been many studies of water flow, sediment trans- This review concerns development of fluvial depositional
port, and channel migration in natural rivers, but not many of models over the past two decades. Earlier studies of fluvial
these have been undertaken over periods of years and including environments are discussed in Miall (1996) and Bridge (2003).
floods (review in Bridge, 2003). Those studies of processes of fluid The presentation in this paper is as follows: (1) definition of
flow, sediment transport, erosion, and deposition in fluvial envi- superimposed scales of fluvial bed forms and associated
ronments during flood stages (e.g., Bridge and Jarvis, 1982; stratasets; and (2) new data and models for each scale of bed
Dietrich and Smith, 1983, 1984; Bridge and Gabel, 1992; Gabel, form and strataset, smallest to largest, with discussion of the use
1993; Julien and Klaasen, 1995; Richardson et al., 1996; Richardson of these models to interpret and predict ancient deposits. The
and Thorne, 1998; Ten Brinke et al., 1999; McLelland et al., 1999; relevance of the new depositional models to exploration and
Ashworth et al., 2000; Julien et al., 2002; Wilbers and Ten Brinke, management of hydrocarbon reservoirs and aquifers is dis-
2003) were facilitated by building bridges over small rivers (less cussed throughout.
than 30 m wide), and use of new equipment such as acoustic
Doppler current profilers, high-resolution multibeam sonar, and SUPERIMPOSED SCALES OF
accurate and rapid positioning using differential GPS. Further FLUVIAL FORMS AND STRATASETS
work of this kind will be necessary in order to make progress in
the development of models that link bed topography, fluid flow, Table 1 and Figure 1 show (in simplified form) the different
sediment transport, erosion, and deposition in rivers and flood- scales of fluvial forms and associated sediment deposits. It will be
plains. demonstrated below that the geometry of a particular type of
Laboratory flumes have been used recently to study relatively strataset is related to the geometry and migration of the associ-
small-scale bed forms (bed-load sheets, ripples, dunes, and ated fluvial form. In particular, the length-to-thickness ratio of
antidunes) and their associated sedimentary structures (e.g., stratasets is similar to the wavelength-to-height ratio of associ-
Bridge and Best, 1988, 1997; Bennett and Bridge, 1995a; Paola et ated forms. Furthermore, the wavelength and height of bed forms
al., 1989; Best and Bridge, 1992; Storms et al., 1999; Alexander et such as dunes and bars are related to channel depth and width.
al., 2001; Leclair and Bridge, 2001; Leclair, 2002); channel and bar Therefore, the thickness of a particular scale of strataset (e.g.,
evolution and migration (e.g., Schumm and Khan, 1972; Ashmore, medium-scale cross sets and large-scale sets of inclined strata)
1982, 1991, 1993; Fujita, 1989; García and Niño, 1993; Whiting and will vary with river dimensions. These relationships between the
Dietrich, 1993a, 1993b; Ashworth, 1996; Lanzoni, 2000a, 2000b; dimensions of stratasets, bed forms, and channels mean that
Knappen et al., 2001); geometry, water flow, and sediment trans- generally applicable, quantitative depositional models can be
port in channels (e.g., Ashmore, 1988; Hoey and Sutherland, 1991; developed.
Warburton and Davies, 1994; Bennett and Bridge, 1995b; Frederici
and Paola, 2003); water flow over simple floodplains (e.g., Knight PLANE BEDS, RIPPLES, DUNES, ANTIDUNES,
and Shiono, 1996; Naish and Sellin, 1996; Sellin and Willetts, 1996; AND ASSOCIATED STRATASETS
Willetts and Rameshwaran, 1996; Wormleaton, 1996; Lai et al.,
2000; Patra and Kar, 2000; Knight and Brown, 2001; Myers et al., The origin, geometry, and migration of (sub–bar scale) bed
2001; Valentine et al., 2001); the effects of vegetation, tectonism, forms such as ripples, dunes, and antidunes, and the nature
base-level change, aggradation, and degradation on rivers and and origin of their associated sedimentary structures, are well
floodplains (or fans) and their deposits (e.g., Ouchi, 1985; Schumm summarized by Allen (1982), Middleton and Southard (1984).
et al., 1987; Germanowski and Schumm, 1993; Leddy et al., 1993; However, there have been some significant improvements in
Koss et al., 1994; Wood et al., 1993; Bryant et al., 1995; Ashworth our knowledge of these bed forms and sedimentary structures
et al., 1999, 2004; Heller et al., 2001; Paola et al., 2001; Gran and over the past two decades (reviews by Best, 1996; Bridge,
Paola, 2001; Moreton et al., 2002; Sheets et al., 2002). Laboratory 2003), including: (1) description of bed forms and their hy-
experiments are desirable in that they are undertaken in manage- draulic stability limits over a broader range of grain sizes than
able environments, and the variables that control the environ- hitherto; (2) description of low-relief bed waves (bed-load
ment can be varied systematically. These physical laboratory sheets) on nominally plane beds, and the realization that their
models may be full scale, reduced scale, or unscaled (analog migration is responsible for the formation of planar strata; (3)
models). However, scaling problems are common, and these analysis of changes in the geometry of bed forms as they
increase as the physical model becomes increasingly smaller than migrate under steady and unsteady flows; (4) detailed exami-
the real-world prototype. Laboratory experimenters have not nation of the way bed-form geometry and migration and
been able to generate all of the superimposed scales of bed forms aggradation rate control the geometry and preservation of
(e.g., bedload sheets, ripples, dunes, unit bars, compound bars) sedimentary structures; and (5) development of quantitative
and associated strata that occur in natural rivers. Also, unrealis- models of the relationship between bed-form height and length
tically high rates of channel migration occur in many experi- and the thickness and length of sets of cross strata. In view of
ments. the fact that these relatively small-scale sedimentary struc-
Quantitative models of fluvial environments can both en- tures can be described from cores and image logs as well as
hance understanding and allow prediction. Because such models outcrops, it is worthwhile investing some time in understand-
are ultimately based on understanding gained from field and ing them.
laboratory studies, the models of short-term, small-scale pro- Dunes are generally recognized as the most common sub–
cesses are best developed. Nevertheless, there have been only bar-scale bed form in sandy rivers. However, contrary to popular
modest advances here, because of the lack of appropriate data perception, dunes also occur commonly in gravelly–sandy rivers
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 87

TABLE 1.—Scales of fluvial forms and associated stratasets (simplified).


Fluvial form Strataset
Ripples Small-scale cross stratasets
Dunes Medium-scale cross stratasets
Low-relief bedwaves on plane beds Planar stratasets
Antidunes Low-angle cross stratasets

Seasonal flood deposit Large-scale (inclined) stratum

Unit bars and compound bars Large-scale inclined stratasets


Channels Channel fills

Channel belt composed of channels and bars Group of large-scale inclined stratasets and
channel fills
Floodplain with levees, crevasse splays, Groups of large-scale inclined stratasets and
channels, lakes, floodbasins channel fills

Alluvial valley with channel belts and floodplain Groups of groups: alluvial architecture
(or fan or delta)

Alluvial river system Basin fill (or part of basin fill)

A 2 B
point bar

channel belt

braid bar
chute
1

sandstone body with sets of large-scale inclined strata: channel belt deposit
2

set of large-scale inclined strata: channel bar deposit


C

large-scale inclined stratum:


seasonal deposit on channel bar
simple: no unit bar compound: unit bar deposit
small-scale cross strata: ripples

planar strata: plane bed

medium-scale cross strata


superimposed on simple
large-scale inclined strata:
medium-scale cross strata: dunes dunes on unit bar

FIG. 1.—Superimposed scales of fluvial forms and stratasets. A) Cross sections (1) and (2) through an idealized braided channel belt.
The cross sections show several sets of large-scale inclined strata formed by deposition on channel bars. Each large-scale inclined
stratum can be simple (deposited during a single flood) or compound (deposited as a unit bar over one or more floods). Large-
scale inclined strata contain smaller-scale stratasets associated with passage of ripples, dunes, and bedload sheets over bars.
Modified from Bridge (2003). B) Alluvial valley of the Senguerr River, southern Argentina, containing a floodplain with a channel
belt (about 100 m wide) on one side of it, adjacent to the valley margin in the foreground. C) Channel-belt sandstones (gray) and
floodplain deposits (red) from the Miocene Siwaliks of northern Pakistan. Marked channel-belt sandstone body is 10–15 m thick.
88 JOHN S. BRIDGE

(e.g., Dinehart, 1989, 1992; Carling, 1999; Kleinhans et al., 2002; formed by dunes is common in pebble-gravelly fluvial deposits
Lunt et al., 2004a, 2004b). One reason why gravel dunes had not (e.g., Lunt et al., 2004a, 2004b), but is perhaps not as easily seen as
been more widely reported is that gravelly–sandy rivers are in sands. Hydraulic stability diagrams for sub–bar-scale bed
rarely studied during peak flood conditions. Also, gravel dunes forms have recently been extended into the gravel sizes (Fig. 2A;
commonly have low heights on the upper parts of bars where Carling, 1999) and the silt sizes (Fig. 2B; van den Berg and van
they can be observed easily at low flow stage. Cross stratification Gelder, 1993).

FIG. 2.—A) Hydraulic conditions for stability of bed forms in sand and gravel, according to Carling (1999) modified from Allen (1982),
as a function of dimensionless bed shear stress and median sediment diameter. Includes additional data from Carling and
Shvidchenko (2002). B) Hydraulic conditions for stability of bed forms in sand and silt, according to van den Berg and van Gelder
(1993), as a function of dimensionless bed shear stress and median sediment diameter.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 89

Planar strata are formed by migration of low-relief bed waves pended sediment load is substantial (Bridge and Best, 1988, 1997;
(bed-load sheets) on lower-stage plane beds and upper-stage Best and Bridge, 1992). The bed forms are millimeters high, and
plane beds (Bennett and Bridge, 1995a; Paola et al., 1989; Bridge their length is about six flow depths. As these bed forms migrate
and Best, 1988, 1997; Best and Bridge, 1992). Thus, plane beds are downstream (at rates of millimeters per second), suspended
not strictly planar. Low-relief bed forms on lower-stage plane sediment is deposited in the trough of the bedform. As the bed
beds are associated with low bed-load transport rates of coarse form migrates over this trough, a lamina is produced that has a
sand to gravel. These bed forms are a few grain diameters high, fine-grained base formed from the suspended-sediment deposi-
and their length is proportional to flow depth (Bennett and tion (Fig. 3C).
Bridge, 1995a). The bed forms are composed of the smaller grain As subcritical bed forms (bed-load sheets, ripples, and dunes)
sizes available in the bed, and the largest grains in the bed are migrate under steady or unsteady flow conditions, the wave-
more or less immobile. The gravel on most of the backs of these length, height, and migration rate of individual bed forms change
bed forms is imbricated, and may be in the form of pebble or in time and space (review by Bridge, 2003, p. 79–97). Modes of
cobble clusters. The steep fronts of the bed forms may have platy change include: formation of small bed forms on the backs of
grains dipping in the flow direction at the angle of repose (pseudo- larger ones; one bed form catching up with the one in front of it,
imbrication). Thus gravelly planar strata formed by downstream thus forming a single bigger bed form; sudden increase in the
migration of low-relief bed waves typically fine upward, and height of a bed form by deep scour of its trough; gradual reduc-
contain both imbricated and pseudo-imbricated gravel clasts tion in height (dying out) of bed forms. As a result of this
(Fig. 3A, B). As bed shear stress increases, grain size of the bed behavior, probability density distributions of bed-form height
load becomes more like the bed material, and low-relief bed and length are normally asymmetrical (high-end tail) and
waves are transformed into dunes. polymodal. This bed-form behavior has an important influence
Bed-load sheets on upper-stage plane beds are associated on the nature and degree of preservation of planar strata and
with high bed-load transport rates of sand mainly, and sus- cross strata (Fig. 4). Changes in the geometry and migration rate

A B

FIG. 3—A) Planar strata in gravel formed by migration of bedload sheets. A planar stratum composed of open-framework gravel
occurs in the middle of the photo. The base of this stratum is relatively coarse grained and contains imbricated pebbles dipping
to the left. The upper part of the stratum is finer grained and contains pseudo-imbricated pebbles dipping to the right. B) bedload
sheets on a bar in the Sagavanirktok River, Alaska. C) Planar strata in sand formed by migration of low-relief bed waves (bedload
sheets) on upper-stage plane beds. Section is 30 mm thick. The dark boundaries of the planar laminae are formed of relatively fine-
grained sediment deposited from suspension in the troughs of bed-load sheets. From experiments of J.S. Bridge and J.L. Best.
90 JOHN S. BRIDGE

FIG. 4.—Cross strata formed by dunes. Flow is right to left and width of section is 0.5 m. Dune form at top of section is marked by
dark line. Arrow 1 indicates thickening of a cross set due to dune trough scouring. Arrow 2 is the front of a dune that was preserved
as the dune was overtaken by another dune. Then, the dune decreased in height and its trough (arrow 3) was filled with low-angle
strata. From experiments of Leclair (2002).

of dunes over floods have received some attention recently (e.g., sm = lr/c + e/a (1)
Gabel, 1993; Julien and Klaassen, 1995; Harbor, 1998; Ten Brinke
et al., 1999; Carling et al., 2000a; Carling et al., 2000b; Julien et al., where sm is mean cross-set thickness, l is mean wavelength of bed
2002; Wilbers and Ten Brinke, 2003), but theoretical models for forms, r is mean deposition rate, c is mean migration rate of bed
such changes are still lacking. forms, e is a parameter that varies with bed-form type from about
Recent experiments on antidune migration under aggrada- 0.8 to 1.6, and a is a parameter that depends on mean bed-form
tional conditions (Alexander et al., 2001) have revealed that the height (commonly about 0.2 mean bed-form height). For dunes
dominant internal structure is trough-shaped stratasets con- and bars, the first term on the right-hand side of Equation 1 is
taining low-angle cross strata (as seen in along-stream sections; normally an order of magnitude less than the second term and
Fig. 5). The cross strata may dip upstream or downstream, or in can be ignored, implying that aggradation rate has little influence
no preferred direction. These stratasets are formed as antidunes on the thickness of stratasets formed by dunes and bars. Aggra-
and associated water-surface waves migrate upstream, increase dation rate can have more of an effect for smaller bed forms such
in height and asymmetry, and then break up, filling the antidune as ripples and low-relief bed forms. Another significant result of
trough with sediment very rapidly. Preservation of antidune the recent studies is that the alongstream length of preserved
forms is rare because they are such short-lived bed forms. stratasets is proportional to the length of the formative bed form
It is commonly stated that aggradation and bed-form climb- (mean strataset length is approximately half mean bedform length
ing are required for preservation of sets of cross strata. In reality, for dunes and antidunes), such that there is a relationship be-
the most important factor controlling cross-set preservation is tween bed-form length/height and strataset length/thickness
the variability of trough scour depth (and height) of the indi- (Fig. 6). This has important implications for predicting the lateral
vidual bed forms. Only those bed forms with the largest scour extent of stratasets from their thickness, but establishment of such
depths leave sets of cross strata, and of these less than a half of relationships for all bed-form scales needs more research.
the height of the bed form is represented as a cross set (Paola and
Borgman, 1991; Best and Bridge, 1992; Bridge, 1997; Bridge and FLOOD-GENERATED STRATASETS
Best, 1997; Storms et al., 1999; Leclair and Bridge, 2001; Leclair,
2002). A simple approximate relationship between the mean The hydraulic conditions controlling the existence of the
thickness of cross sets, the mean height of bed forms, and the various bed forms discussed above (Fig. 2) are fairly well known
aggradation rate is as long as the bed forms are in equilibrium with the flow (reviews

antidune

FIG. 5.—Cross strata formed by antidunes. Flow is right to left, and section is 1 m wide. Antidune (marked) occurs at top of section.
Cross strata formed by aggradation on migrating antidunes are inclined at a low angle and fill troughs. Relatively coarse sediment
is light, and fine sediment is dark. Troughs are lined with fine-grained sediment. From experiments of Alexander et al. (2001).
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 91

FIG. 6.—Scales of bedforms and associated stratasets from modern rivers. From Bridge and Lunt (2006).

by Allen, 1982; Bridge, 2003). This information is useful because repose) than their stoss side (Fig. 8A). Point bars and braid bars
it allows prediction of changes in bed forms (hence sedimentary are normally compound bars in that they are composed of unit
structures) as flow conditions change in time and space over bars that accreted onto the compound bars during floods. Accre-
sediment beds of a given grain size. However, changes in the tion of unit bars is indicated by accretion topography in the form
geometry of large bed forms such as dunes normally lag behind of bar-head lobes and bar-tail scrolls (Fig. 8). Unit bars are not
changes in flow conditions, because of the amount of sediment always obvious in rivers at low flow stage, because they may be
that must be moved to effect a change in geometry. This means low-relief features, or they may be modified during falling flow
that prediction of the flow conditions for dune existence under stage. Cross-bar channels (including chute channels) commonly
unsteady flows is imprecise. Nevertheless, Figure 7 shows some cut through compound bars, especially between individual unit
typical sedimentary sequences in which mean grain size and bars (Fig. 8).
sedimentary structures vary vertically and laterally, associated The geometry, spatial distribution, and migration of bars
with changing flow conditions over a typical flood. Temporal within channels control the plan geometry (channel pattern) of
changes in sediment transport rate and grain size (hence bed the channel belt: that is, the sinuosity of channels and the
form) at a point are associated with changes in flow strength (e.g., degree of channel splitting (braiding). These are in turn con-
flow velocity or bed shear stress) over the flood, and these trolled mainly by the supply of water and sediment during
produce vertical variations in grain size and sedimentary struc- floods (channel-forming discharge). A continuum of channel
tures. Spatial decrease in sediment transport rate is mainly re- patterns occurs as water supply and sediment supply are
sponsible for the deposition and for downstream changes in increased (Fig. 9): single (meandering) channel of increasing
mean grain size and sedimentary structures. The availability of sinuosity; single to braided channel; braided channel with
different sediment sizes and the overall flow strength controls the increasing degree of braiding. All of these different channel
overall grain size of the deposited sediment. Sequences such as patterns can be formed at constant discharge, and in muddy,
those shown in Figure 7 will be incorporated into larger-scale sandy, and gravelly rivers. However, their detailed geometry
sequences associated with channel bars and channel fills, levees, is influenced by discharge variations, sediment size supplied,
and crevasse splays. and riparian vegetation.

RIVER CHANNELS, BARS, AND CHANNEL BELTS Water Flow.—

General Character of Geometry, Water Flow, Sediment Water flow in single curved channels is broadly equivalent to
Transport, Erosion, and Deposition the flow in curved channels around braid bars. Curved flow
around and over any type of channel bar results in: (1) a cross-
Channel Geometry.— stream component of water-surface slope towards the center of
curvature; (2) a cross-stream (secondary) flow pattern; and (3)
The fundamental components of all alluvial channel belts are convective acceleration and deceleration of the depth-averaged
channels, unit bars, and compound bars (Fig. 8). A bar is defined downstream flow associated with bar topography (Fig. 10). As a
as a bed form with length proportional to local channel width and result, the maximum depth-averaged flow velocity moves from
height proportional to channel depth. Unit bars are lobate the convex bank at the bend entrance (the shallow bar-head
(linguoid) in plan, and their lee side is steeper (up to the angle of region) towards the concave bank (the deep region adjacent to the
92 JOHN S. BRIDGE

FIG. 7.—Typical sedimentary sequences produced by erosion and deposition over a single flood period, for three different grain-size
ranges. Deposition, and downstream change in grain size and sedimentary structures, are caused by decreasing bed shear stress
and sediment transport rate (i) in the flow direction (x). Vertical changes in grain size and sedimentary structures are caused by
change in bed shear stress and sediment transport rate with time (t) over the flood period. Thickness of flood-generated
sedimentary sequences is typically centimeters to meters.

A B

a
b c
b

FIG. 8.—A) Geometry of unit bars, compound bars, and channels in a braided channel belt, Rakaia River, New Zealand. Photo from
Jim Best. Flow is right to left, and width of view is about 500 m. Single lobate unit bars occur in the channel at (a). An incipient
compound braid bar occurs at (b). The bar head is composed of a single, partly emergent lobate unit bar that is dissected by a series
of cross-bar channels. These cross-bar channels have small solitary bars (deltas) at their downstream ends. The bar tail is composed
of two emergent scroll bars and at least one incipient submerged scroll bar. A well-developed compound braid bar occurs at (c).
The bar head is composed of at least six lobate unit bars, and the bar tail is composed mainly of a single scroll bar. Cross-bar
channels are also evident. B) Geometry of unit bars, compound bars, and channels in a meandering channel belt. Flow is left to
right. Compound point bar in center of photo contains lobate unit bars at the bar head (a), scroll bars at the bar tail (b), and
abandoned cross-bar channels (c). Madison River near Hebgen Lake, Montana, U.S.A. Channel is about 50 m wide.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 93

bar tail) with progression around the bend. The cross-stream


components of water-surface slope and flow velocity increase
relative to the downstream components as the radius of bend
curvature decreases. In general, the mean cross-stream flow veloc-
ity is an order of magnitude less than the mean downstream flow
velocity. Bed shear stress varies in a way similar to that of depth-
averaged flow velocity. However, this is a simplified view of flow
in curved channels, and more details are given in Bridge (2003).
Water flow in channel confluences is somewhat analogous
to that in adjacent curved channels if the joining channels have
similar geometry (Fig. 10). The maximum high-stage velocity
occurs near the center of the confluence scour, with lower
velocities near the banks (Fig. 10). However, the pattern of flow
in confluences is also dependent upon flow acceleration associ-
ated with reduced cross-sectional area of the conjoined streams,
flow separation downstream from crests of tributary mouth-
bars, inequality in the depths of confluent channels, and the
FIG. 9.—Continuum of channel patterns. Channel sinuosity and enhanced turbulence of the mixing layer between the joining
degree of channel splitting around braid bars vary with streams (Fig. 10).
channel-forming (bankfull) water discharge, valley slope, Water flow in channel diffluences (upstream of braid-bar
and mean grain size of transported sediment. Valley slope is heads) is not known well, because of lack of data (review in
a surrogate measure of sediment transport rate. Bridge, 2003, p. 188; see also Frederici and Paola, 2003). At high
A
A B
A
C
B
D
A
C

E
D

Flow

B B
F G

Flow

Mouth bar
G

FIG. 10.—A) Simplified pattern of near-bankfull water flow for curved channels of similar geometry adjacent to a braid bar (sections
A, B, C), and the downstream confluence (section D) and diffluence (section E) regions. For simplicity, flow patterns associated
with unit bars and smaller bed forms are not included. Arrows on the map represent vectors of depth-averaged, downstream flow
velocity. The cross sections show channel geometry and flow-velocity vectors (near surface and bed) for the cross-stream flow.
Cross-stream flow velocities are typically an order of magnitude less than downstream flow velocities. These flow patterns and
channel geometry change with flow stage. B) Simplified pattern of near-bankfull flow for a confluence in which a relatively small
channel with a tributary mouth bar joins a larger channel. Symbols as for Part A. From Bridge (1993), and based on the work of
Jim Best and Andre Roy.
94 JOHN S. BRIDGE

flow stage, the locus of maximum velocity is in midchannel the bankfull flow pattern in the main channel (review in Bridge,
upstream of the diffluence, and splits downstream such that 2003).
each high-velocity thread is close to the upstream tip of the braid
bar. It is common for the relative discharges of the split channels Sediment Transport and Bed Forms.—
to vary with time, and one of the channels may become domi-
nant while the other fills. Diffluence zones may also contain Bed-load transport rate and mean grain size at channel-
complicated patterns of convergence and divergence of depth- forming flow stage over point bars and braid bars generally
averaged flow velocity. increase with depth-averaged flow velocity and bed shear stress
The flow patterns described above for near-bankfull (chan- (Fig. 11), which are largest in the shallow water near the up-
nel-forming) flow stage change with water discharge. At rela- stream end of a bar, in mid-channel in the mid-bar region, and
tively low discharge, depth-averaged flow velocity and bed shear in the deep water adjacent to the bar tail. Bed material normally
stress are less than for high discharge. Water flows in a more fines downstream on the tops of both braid bars and point bars
sinuous path around emerging bars, resulting in relatively strong in modern sandy and gravelly rivers, and is relatively coarse in
cross-stream flow components relative to downstream flow com- the deepest parts adjacent to the bar tail. Thus, the spatial
ponents. Dissection of emerging bed forms such as dunes and distribution of mean grain size of the bed material reflects the
unit bars may result in small cross-bar channels. At high dis- distribution of bed-load grain size and bed shear stress at high
charge, new channels may be cut across existing compound bars, (channel-forming) discharges, as would be expected. In
particularly through the low areas between adjacent unit bars, confluences, the largest bed-load transport rates also generally
and the relative discharges of braided channels may vary. During occur where the flow velocities and bed shear stresses are
overbank floods, the overbank flow may interact with and modify largest, provided that the bed is not armored. The largest mean

A
A
p les
rip
rip
upper-stage plane beds and p les
antidunes (shallow, fast flow)
rip Fine Flow
p les les
(sh ripp
a llow Coarse
, sl Dunes and bed-load sheets
ow
flow over most of bed
)

FIG. 11.—A) Schematic distribution of bedload grain size and sub–bar-scale bed forms in sandy and gravelly rivers at bankfull flow
stage. Ripples occur only in sands with diameter less than about 0.7 mm. The boundary between coarse and fine sediment is
actually gradational. B) Dunes preserved on the upper part of a point bar (Congaree River, South Carolina, U.S.A.) following a
flood. Medium-scale trough cross strata exposed in trench in foreground. The scale in the trench is 0.15 m long, and the trench
is about 0.75 m deep.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 95

grain size occurs in the deepest, downstream part of the The variation of sediment transport rate and bed sediment
confluence scour, whereas the finest mean grain size occurs size in alluvial channels is reflected in the bed forms superim-
immediately upstream of the scour zone and near the banks posed upon bars. Dunes with curved crest lines are the most
adjacent to the downstream end of the confluence (Fig. 11). common bed forms in sandy rivers at high flow stage (Fig. 11).
These patterns of bed-load transport rate and mean grain size Upper-stage plane beds in sands occur locally in shallow areas of
would be somewhat different if the braided channels differed high flow velocity. Ripples can occur only in sands with mean
greatly in discharge and geometry. diameter less than about 0.7 mm, and are normally restricted to
As discharge changes, the patterns of bed topography, flow areas of slow-moving water near banks. Dunes are also common
velocity, and bed shear stress also change, as discussed above. in gravelly–sandy rivers at high flow stage, as are lower-stage
The decreased bed shear stress at low flow stage generally results plane beds (with bed-load sheets, pebble clusters, and sand
in smaller transport rates and grain sizes of bed load. Bed-load ribbons). Antidunes in sands and their equivalent in gravels
transport may essentially cease at low flow stage in areas with (transverse ribs) occur only rarely in fast, shallow water. At flow
gravelly bed sediment, and armor layers may develop. Changes stages lower than bankfull, dunes are generally shorter and
in the spatial distribution of bed shear stress with changing flow lower, and the proportion of curved-crested dunes decreases
stage may bring a zone of high bed shear stress over a zone of relative to ripples and lower-stage plane beds. Dune geometry is
relatively fine-grained bed material, thus producing high bed- commonly not in equilibrium with rapidly changing flow stage,
load transport rate locally. This typically happens adjacent to bar particularly in shallow water, where dunes can become exposed
tails. by small decreases in water level.
In general, suspended-sediment concentrations are relatively
large where bed shear stress and turbulence intensity are large Erosion and Deposition.—
and where bed-material size is small. Suspended-sediment con-
centrations are very difficult to predict in confluence zones in Erosion and deposition at the scale of the channel bars and
view of the zones of mixing, upwelling, and flow separation. bend scales are due to: (1) adjustments of bed topography specifi-

A
B

Flow

Potential rising-stage erosion, falling-stage deposition


Potential cross-bar channel
Potential rising-stage deposition, falling-stage erosion directions at high stage

DOWNSTREAM PART OF CURVED CHANNEL


falling-stage erosion
A
3 2
falling-stage deposit 1
4 rising-stage deposit

flood-stage deposit

CURVED CHANNEL ENTRANCE

falling-stage deposit
B flood-stage deposit
falling-stage deposit

2 3
1 4
rising-stage deposit flood-stage deposit

FIG. 12.—Typical pattern of erosion and deposition at the channel bar and bend scale for the case of a symmetrical braid bar (modified
from Bridge, 1993). Cross sections show channel geometry during the course of a flood: (1) pre-flood low stage; (2) flood stage;
(3) flood stage after bank erosion and bar deposition; (4) post-flood low stage.
96 JOHN S. BRIDGE

cally associated with varying discharge (Fig. 12); (2) bank erosion point bars (Fig. 13). Another well-known type of bend cutoff,
and associated bar deposition, occurring mainly at high flow neck cutoff, occurs in very sinuous rivers (Fig. 13), and is enter-
stage; and (3) cutting of new channels, enlarging existing chan- tainingly discussed by Mark Twain in Life on the Mississippi.
nels, and filling of others. During rising flow stage, erosion tends
to occur in the deepest parts of bends and confluence scours and Misconceptions About River Channels
the upstream ends of bars, whereas these areas receive deposits
during falling stages (Fig. 12). In contrast, the highest parts of bar Several misconceptions about the nature of alluvial river
tails tend to be areas of deposition at high flow stage, with erosion channels that occur in much of the literature need to be dispelled
as stage falls. Such adjustments in bed topography are normally (Bridge, 2003). First of all, river channel patterns cannot be
associated with bank erosion and deposition on adjacent bar properly classified using the terms straight, meandering, braided,
margins. Bar migration during high flow stages is most com- and anastomosing (e.g., Miall, 1996). Straight alluvial channels
monly by lateral and downstream accretion (Fig. 13). Such chan- occur only where the flow is not powerful enough to erode the
nel accretion is episodic, and may be in the form of distinct unit channel banks. This may be because of high bank resistance
bars. The upstream ends of channel bars are sites of erosion (caused by early lithification), but normally because of human
during bankfull flow stage but may receive deposits at high engineering. Anastomosing rivers do not belong in a classifica-
falling stages. Erosional enlargement of one braided channel tion based on channel sinuosity and degree of splitting around
while an adjacent channel is filled is commonly associated with braid bars. Anastomosing channels are divided by areas of flood-
migration of unit bars into a channel entrance, thereby blocking plain and are long enough to contain many bars, enabling assign-
discharge into the channel. The enlarging channel continues to ment of channel pattern based on degree of channel splitting
migrate laterally and downstream while the blocked channel fills around bars and sinuosity. This means that the terms anastomos-
with sediment, resulting into an asymmetrical bar form in plan. ing and braiding are not mutually exclusive, and cannot be used
The growing bar tail of the enlarging channel obstructs the together in a single classification. In fact, many braided rivers
downstream end of the diminishing channel (Fig. 13; Ashworth appear to be both braided and anastomosing (Fig. 14). The term
et al., 2000; Bridge, 2003). New channels are commonly cut across anastomosing belongs in a classification of patterns of channel-belt
low areas adjacent to unit bars at high and falling flow stage, branching and joining, a classification that also includes the terms
leading to chute cutoff of bends and formation of braid bars from tributive and distributive.

Translation
Translation and expansion

Translation, expansion, and


development of meander lobes

Chute cutoff
Neck cutoff

FIG. 13.—A) Typical modes of channel migration for single-channel rivers (from Bridge, 2003). Active channels are stippled.
Simplified accretionary units on point bars (separated by lines) are actually composed of unit bars (bar-head lobes and bar-tail
scrolls). Upper photo (courtesy of Jim Best) shows accretion topography on a Paraná River (Argentina) point bar adjacent to an
abandoning channel. Lower photo shows neck cutoff of a meander loop of the Calamus River, Nebraska, USA. Channel width
is about 15 m.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 97

Another common misconception is that discharge variability lowlands (e.g., the U.S. Great Plains). In fact, many braided rivers
is greater for braided rivers than for single-channel rivers. This carry large amounts of sand and silt in suspension (e.g.,
misconception probably originated from the early studies of Brahmaputra in Bangladesh, Yellow in China, Platte in Ne-
proglacial braided rivers in mountainous regions of North braska), and many single-channel, sinuous rivers carry sands and
America, where discharge varied tremendously during snow- gravels as bed load (Madison in Montana, South Esk in Scotland,
melt. In contrast, many single-channel rivers were studied in Yukon in Alaska).
temperate lowland regions where discharge variations were Linked to the perceived lateral instability of braided rivers
moderated by groundwater supply. In fact, discharge variability relative to meandering rivers, plus relatively large channel width/
does not have a major influence on the existence of the different depth of braided channels, is the supposition that the width/
channel patterns, because they can all be formed in laboratory thickness of braided channel belts is much greater than meander-
channels at constant discharge, and many rivers with a given ing channel belts. This is not generally the case. Channel belts of
discharge regime show along-stream variations in channel pat- meandering Mississippi and braided Brahmaputra have similar
tern. width/thickness (Mississippi: 15 km / 40 m = 375; Brahmaputra:
Another misconception is that rivers that transport large 10 km / 40 m = 250).
amounts of bed load relative to suspended load have relatively The influence of vegetation on alluvial rivers is also misun-
low sinuosity and high degree of braiding (Schumm, 1977; Miall, derstood. Vegetation helps stabilize channel banks and bar
1996). Such bedload streams have been associated with relatively surfaces given adequate time and conditions for development.
easily eroded banks of sand and gravel, large channel slope, and Such bank stabilization allows the existence of relatively steep
large stream power, such that they are laterally unstable. In con- cut banks and may hinder lateral migration of channels. It has
trast, rivers with relatively large suspended loads were postu- been claimed that all rivers prior to the development of vegeta-
lated to be characteristic of undivided rivers of higher sinuosity. tion on land were braided because of abundant supplies of
Such suspended-load streams were associated with cohesive muddy relatively coarse sediment and their banks were “unstable”
banks, low stream gradient and power, and lateral stability. This (Schumm, 1968; Cotter, 1978). However, there is no conclusive
misconception probably arose because early studies of braided evidence that vegetation (or early lithification) has a significant
rivers were in mountainous areas of sandy–gravelly outwash and influence on channel pattern, as long as the flood flow is capable
those of single-channel sinuous streams were from temperate of eroding banks and transporting sediment. Low-powered

B Translation with symmetrical channels

Translation and expansion with symmetrical channels

Translation and changes in relative discharge of channels

FIG. 13 (continued).—B) Typical modes of channel migration for simple braided-river patterns (from Bridge, 2003). Simplified
accretionary units on braid bars (separated by lines) are actually composed of unit bars (bar-head lobes and bar-tail scrolls). The
braid bar grows asymmetrically in the case where discharge in one channel is increasing at the expense of the other channel. Photo
(Sagavanirktok River, northern Alaska, U.S.A.) shows braid bar with accretion topography, active channel to left and filling
channel to right. Filling channel (about 120 m wide) contains lobate unit bars.
98 JOHN S. BRIDGE

Point bar

Meandering and braided channel


Braid bar

Bar assemblage
(floodplain)

Braided channel

Floodplain (abandoned-
bar assemblages)

FIG. 14.—Sagavanirktok (northern Alaska, U.S.A.) channel belt with compound braid bars and point bars associated with braided
and meandering channels, and anastomosing, braided–meandering channels separated by a bar assemblage. Compound bars
have accretion topography indicating downstream translation and lateral growth, and channel fills are also evident. Channel
belt is 2 km wide.

streams may not be capable of eroding banks and transporting bars to increase in width (Fig. 13). In some cases, there is even
appreciable amounts of sediment, thus allowing vegetation to upstream accretion of bars. Thus, the key differences between
encroach into the channel (e.g., streams in lowland swamps). meandering rivers and braided rivers are the more frequent
occurrence in braided rivers of braid bars bounded by coeval
Similarities and Differences Among channels, and of confluence regions bounded by coeval side bars
Different Channel Patterns (Fig. 8).
Deposition on both point bars and braid bars is commonly in
The main types of channel pattern are single-channel (mean- the form of unit-bar accretion during floods. Unit bars and
dering) and braided. Both types of channel pattern have curved supposedly related sets of planar cross strata have been specifi-
channel segments adjacent to compound bars. Braided channels cally associated with braided rivers (e.g., Collinson, 1970; Smith,
have zones where the flow divides and rejoins around compound 1971, 1972, 1974; Bluck, 1976,1979; Cant, 1978; Cant and Walker,
braid bars; however, cutoff of point bars in meandering rivers is 1976,1978; Blodgett and Stanley, 1980; Crowley, 1983). However,
also a braiding pattern. The patterns of flow and sediment trans- unit bars occur in meandering rivers also (e.g., McGowen and
port around curved channel segments are the same irrespective Garner, 1970; Bluck, 1971; Jackson, 1976; Levey, 1978; Bridge et al.,
of whether the bar is a point bar or a braid bar. All river channels 1995). Furthermore, most of the internal structure of unit bars is
migrate by erosion of concave banks in curved channel segments not planar cross strata, but is due to the bed forms (dunes, ripples,
and deposition on compound bars. Deposition is mainly on the bedload sheets) migrating over them (Collinson, 1970; Jackson,
downstream parts of point bars and braid bars, but there is also 1976; Nanson, 1980; Bridge et al., 1986; Bridge et al., 1995; Bridge
commonly a lateral component that causes point bars and braid et al., 1998; Ashworth et al., 2000; Best et al., 2003; Lunt et al.,
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 99

2004a, 2004b). Angle-of-repose (planar) cross strata are appar- as negligible over the time spans considered in the models.
ently restricted to the margins of unit bars, particularly where the Despite being simplified, these models give important insights
bars migrate into relatively deep, slow-moving water. into the nature of channel bar deposits that did not come from
The bed forms superimposed on bars during floods are most earlier static 1-D and 2-D models (review in Bridge, 2003). Ex-
commonly dunes, irrespective of grain size or channel pattern amples of these models shown in Figures 15 and 16 illustrate a
(Fig. 11). Ripples occur mainly in very fine to medium sands in number of fundamentally important aspects of river channel
areas of slow flow near banks. Upper-stage plane beds and deposits:
antidunes occur in limited areas of fast shallow flow near bar
tops. However, there are more bedload sheets (planar strata) in 1 As channels migrate by lateral and downstream migration,
gravelly–sandy rivers and more upper-stage plane beds (planar the deposits from different parts of channel bars become
laminae) and ripples (small-scale cross strata) in sandy rivers. vertically superimposed (e.g., bar-head deposits overlying
bar-tail deposits, bar-tail deposits overlying confluence scour
Quantitative Process-Based Models for River Channels deposits).

Quantitative, dynamic, 3-D depositional models of river chan- 2 Systematic spatial variations in the thickness of channel de-
nel deposits have been developed only recently, and such models posits, and the inclination and orientation of large-scale strata,
are at a rudimentary stage (e.g., Willis, 1989; Bridge, 1993). These are due to bed topography and the mode of channel migra-
types of models require prediction of the interaction between bed tion. For example, it is common for channel-bar deposits to
topography, water flow, sediment transport rate, mean grain size thicken (by up to a factor of two), and for large-scale strata to
of bedload, and bed forms within channels of prescribed geom- steepen, towards a cut bank (channel-belt margin) or
etry. The flow conditions are assumed to be steady and bankfull, confluence scour.
with the bed topography, water flow, and sediment transport in
equilibrium. The models apply to either single channel bends 3 Lateral and vertical variation in grain size and sedimentary
with an associated point bar, or two channel bends separated by structures are controlled by the bed topography, flow, sedi-
a braid bar. The plan forms of the channels are sine-generated ment transport and bed forms, and by the mode of channel
curves, and features such as unit bars and cross-bar channels are migration. Channel-bar deposits normally fine upwards, but
not considered. The channels must be put in a dynamic context by they also commonly show little vertical variation in grain size.
allowing them to migrate by bank erosion and bar deposition, Some channel-bar deposits coarsen at the top if bar-head
and to change geometry in time. Net vertical deposition is taken deposits are preserved.
METERS

METERS
METERS

METERS

FIG. 15.—Example of quantitative model of point-bar deposits (Willis, 1989). Meander plans to right indicate downstream and lateral
growth of a point bar in discrete increments, and position of cross sections in various orientations. Cross sections indicate basal
erosion surface of point-bar deposits, large-scale inclined strata due to incremental deposition, contours of mean grain size (dotted
lines annotated in millimeters), and current orientations relative to the cross section (arrows pointing down indicate flow out of
plane of cross section). Point-bar deposits thicken, and inclination of large-scale inclined strata increases, from left to right in
sections A, B, and C.
100 JOHN S. BRIDGE

c1 c5

10 m

0 m 2 section c1
0
m
1

section c5

fill fill
0.5

0 0.5 mm

FIG. 16.—Example of quantitative model of braid-bar deposits (Bridge, 1993). Upper figure shows plan geometry of initial braided
channels (stippled) and migrated channels (dashed). The braid bar migrated downstream in four discrete increments. Cross
sections show basal erosion surface of bar deposits, large-scale inclined strata due to incremental deposition, and details of
spatial variation in deposit thickness, grain size, sedimentary structure, and paleocurrents. Deposit thickness and inclination
of large-scale inclined strata vary systematically. Bar sequences generally either fine upwards or have little vertical variation
in grain size. The dominant internal structure in this example is medium-scale trough cross strata (formed by dunes), with
subordinate small-scale cross strata (formed by ripples).

These models of Bridge and Willis predict the geometry, grain and bed topography. Quantitative models for the flow, sediment
size, and sedimentary structure of the deposits of single point transport, and deposition in abandoned channel fills are also
bars or braid bars. However, they do not consider the somewhat needed.
complicated flow structures at channel diffluences and
confluences. It is necessary to develop theoretical models for flow Qualitative Depositional Models for River Channels
and sediment transport in these regions. Although there are
numerical models of turbulent flow in confluences that agree The qualitative depositional models shown in Figures 17 to 21
fairly well with observed flows (review in Bridge, 2003), they do comprise: (1) maps showing idealized active and abandoned
not describe the interaction between flow, sediment transport, channels, compound bars, and lobate unit bars; (2) cross sections
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 101

F
flow D flow
100s to 1000s m

B
G
Old braid bar

C F
flow
E
100s to
1000s m

flow

FIG. 17.—Qualitative models of channel deposits. Maps of meandering and braided rivers showing active and abandoned channels,
compound bars, and some lobate unit bars. Stippled areas are filled with water at low flow stage, and unit bars within these areas
are not shown. Bar heads of compound braid bars and point bars have formed by accretion of the fronts of lobate unit bars, and
their bar tails have formed by accretion of the sides of lobate unit bars (i.e., scroll bars). Thus, growth of compound bars has been
mainly by lateral and downstream accretion. The upstream end of the abandoned braid channel was blocked by a compound
point bar, and the channel is being filled with unit-bar deposits. The upstream and downstream ends of the abandoned
meandering channel were also blocked by bar deposits, but unit bars in the filling channel do not fill the entire abandoned channel.
Cross sections and vertical sedimentary logs are shown in Figures 18 to 21. Cross sections with letters correspond to those in Figure
18.

showing large-scale inclined strata and their internal structures, channels and braided channels. The bar-head regions of the
associated with migration of compound bars, unit bars, and their compound bars have formed by accretion of the fronts of lobate
superimposed bed forms; and (3) vertical logs of typical sedimen- unit bars, and their bar-tail regions have formed by accretion of
tary sequences through different parts of compound-bar deposits the sides of lobate unit bars (i.e., scroll bars) (Fig. 17). Thus,
and channel fills. The cross sections and vertical logs differ growth and migration of compound bars have been mainly by
somewhat between gravelly–sandy rivers (Figs. 18, 19) and sandy lateral and downstream accretion. The abandoned channel in the
rivers (Figs. 20, 21), and the cross sections differ between single braided-channel model is being filled with unit-bar deposits (Fig.
102 JOHN S. BRIDGE

A Across-stream view of compound side bars adjacent to a confluence scour

lobate unit bar

side bar
side bar
confluence scour

B Across-stream view of compound braid bar that migrated over a confluence

Cross-bar channel
Unit bar

cross-stratified sand in small channel fill

medium-scale cross-stratified sandy gravel medium-scale cross stratified open-framework gravel

C Across-stream view of compound point bar that accreted laterally

D Along-stream view through compound bar that migrated laterally and downstream

Downstream Cross-bar channel Upstream

FIG. 18.—Qualitative models of channel deposits. Cross sections (letters correspond to those in Figure 17) showing large-scale
inclined strata, and their internal structures, associated with migration of compound bars, unit bars, and their superimposed
bedforms for gravelly–sandy rivers. Cross sections are hundreds of meters to kilometers wide and meters thick. Vertical
exaggerations are 5 to 10.

17), and its upstream end was blocked by a compound point bar. others show convex-upward or concave-upward stratal inclina-
The abandoned channel in the meandering-channel model (Fig. tions. The deposits of braided and nonbraided rivers can be
17) is filled with unit-bar deposits only at the upstream end, and distinguished on the basis of these patterns of large-scale inclined
the downstream end is a lake. strata in cross section. This distinction between channel patterns
cannot be made from vertical lithofacies profiles, contrary to
Channel-Bar Deposits.— published opinions. The definitive depositional evidence for
braiding in ancient deposits is cross sections through braid bars
Figures 17, 18, and 20 show how the geometry and mode of with adjacent, coeval channels, and confluences. Examples of
migration of river channels and bars control the geometry and these patterns of large-scale inclined strata from modern channel
orientation of large-scale inclined strata. Large-scale strataset (story) belts are shown in Figure 22.
thickness in a channel belt can vary laterally by a factor of two or Large-scale inclined strata shown in Figures 18 and 20 rarely
more. In places, large-scale stratasets thicken laterally as the have such systematic inclinations, and both discontinuities and
large-scale strata increase in inclination. Some sets have large- discordances are common. Discontinuities in inclination may be
scale strata inclined predominantly in one direction, whereas associated with the occurrence of unit bars (discussed below) and
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 103

E Along-stream view through upstream end of large channel fill: lateral and downstream growth of compound bar

Upstream unit bar deposit Downstream

medium-scale cross stratified open-framework gravel

F Across-stream view of upstream end of large channel fill: lateral accretion and channel filling

G Across-stream view of downstream end of large channel fill: scroll bar accretion and channel filling
scroll bars

FIG. 18 (continued).—

lower-bar platforms (Fig. 22). Discordances in large-scale in- mean grain size in such sequences increases with channel sinuos-
clined strata form through discharge fluctuations and shifts in ity. Bar sequences with little vertical variation in mean grain size
channel position, and are (as discussed below) related to the occur where bend-apex deposits build over bar tails. Such se-
formation of cross-bar channels. quences may coarsen at the top if the bar head migrates over bar-
The number of large-scale inclined strata comprising most of tail deposits. Bar-head deposits can be preserved only if the
the thickness of a set (story) is commonly between 1 and 10, upstream part of the bar is not extensively eroded. Thus, different
dependent on the rate of channel migration relative to channel- types of vertical sequence of lithofacies depend mainly on the
bar width. For example, if the channel migrates a distance position in the bar and on the mode of channel migration rather
equivalent to the apparent bar width during a single deposi- than on channel pattern. In some channel deposits, upper-bar
tional event, the bar sequence comprises a single large-scale deposits and lower-bar deposits can be distinguished by their
stratum. If ten depositional episodes are required to migrate one differences in grain size and sedimentary structure, and by the
bar width, ten large-scale strata are formed. The amount of more common presence of buried vegetation in upper-bar depos-
channel migration during a flood is commonly on the order of its (Figs. 21, 22). Upper-bar deposits commonly increase in thick-
10-1 x channel width. ness in the down-bar direction, whereas lower-bar deposits de-
Downstream translation of bars results in preferential preser- crease in thickness.
vation of bar-tail deposits and erosion of bar-head deposits. Bar- Individual large-scale inclined strata can be recognized by
tail deposits fine upward (Figs. 19, 21), and the vertical range of vertical changes in grain size and sedimentary structure: they
104 JOHN S. BRIDGE

Downstream end
of channel bar
Upstream end
of channel bar
Upstream end
of channel bar

sand gravel sand gravel sand gravel

Channel fill LEGEND


small-scale
cross sets
compound planar strata
large-scale simple
metres set large-scale
set medium-scale
cross sets

sand gravel
vfs fs ms cs vcs vfp fp mp cp vcp
4 3 2 1 0 -1 -2 -3 -4 -5 -6
Grain size (phi)

FIG. 19.—Qualitative models of channel deposits. Vertical logs of typical sedimentary sequences through different parts of compound
bar deposits and channel fills for gravelly–sandy rivers.

commonly fine upwards at the top (Figs. 19, 21, 22). The internal accretion of the deposits of the smaller-scale bedforms that mi-
structure of large-scale inclined strata in sandy and gravelly grate over them: bedload sheets, dunes, or ripples. During their
braid bars and point bars is normally dominated by medium- growth and migration, such unit bars are asymmetrical in along-
scale trough cross strata, by virtue of the ubiquitous presence of stream profile with a relatively steep downstream side that is less
curved-crested dunes on bar surfaces during high flow stage. than the angle of repose. Therefore, cross-set bases and planar
Sets of medium-scale cross strata are commonly decimeters strata tend to be inclined at a relatively low angle (up to about
thick—that is, an order of magnitude thinner than channel-bar 10°), reflecting the geometry of the unit bar over which they are
thickness. Gravelly–sandy deposits may have relatively more migrating. As a unit bar grows, the lee face may reach the angle
planar strata with imbricated pebbles or cobbles, which are of repose, and superimposed bedforms then halt at the crest of the
formed by migration of bedload sheets on lower-stage plane unit bar, from where their sediment avalanches. Therefore, the
beds. Planar strata associated with upper-stage plane beds are low-angle inclined set boundaries within unit bars pass laterally
common in the upper parts of sandy braid bars and point bars. into angle-of-repose cross strata, defining a smaller scale of large-
Small-scale cross-stratification from ripple migration is restricted scale inclined strata than associated with the compound bar on
to relatively fine-grained sands deposited near banks at high which the unit bar is superimposed (Figs. 18, 20, 23). The angle-
flow stage, or in other positions in low-stage deposits. Small- of-repose cross strata associated with unit bars are commonly
scale cross-stratified, bioturbated sand commonly occurs inter- referred to as planar cross strata, and are thought to be character-
bedded with vegetation-rich mud as centimeter-thick units in istic of braided rivers (see above). In reality, unit bars and their
the upper-bar deposits near channel banks. Antidune cross deposits form in all river types, angle-of-repose cross stratifica-
stratification and transverse ribs occur rarely in the upper parts tion is not normally the dominant internal sedimentary structure,
of sandy and gravelly channel bars, that is, where flow is fast and such cross strata look planar only in sections that are small
and shallow. relative to the unit bar. Angle-of-repose cross strata formed by
Large-scale inclined strata may be sheet-like or, if associated unit bars can easily be confused with that due to dune migration.
with deposition of unit bars, mound-like (Figs. 17, 18, 20). Unit Unit-bar deposits associated with bar-head lobes and bar-tail
bars commonly grow in height and migrate downstream by scrolls tend to occur in the upper parts of compound-bar deposits,
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 105

ALONG-STREAM SECTION ALONG AXIS OF COMPOUND BAR

Upstream Downstream
or erosional cross-bar channel fill
channel margin scroll bars

small-scale cross strata

planar strata unit-bar deposit


medium-scale cross strata

ACROSS-STREAM SECTION OF TAIL OF COMPOUND BRAID BAR THAT MIGRATED DOWNSTREAM OVER
CONFLUENCE SCOUR
channel fill braid-bar tail tributary-mouth bars

ACROSS-STREAM SECTION OF CONFLUENCE SCOUR AND ADJACENT SIDE BARS


THAT MIGRATED DOWNSTREAM OVER A BRAID BAR
confluence scour fill side bar tail

ACROSS-STREAM SECTION OF DOWNSTREAM END OF CHANNEL FILL

FIG. 20.—Qualitative models of channel deposits. Cross sections showing large-scale inclined strata, and their internal structures,
associated with migration of compound bars, unit bars, and their superimposed bedforms for sandy rivers. Cross sections are
hundreds of meters to kilometers across and meters to tens of meters thick. Vertical exaggerations are approximately 5.
106 JOHN S. BRIDGE

DOWNSTREAM PARTS OF CHANNEL BARS

BANKFULL LEVEL

Upper bar
Upper bar

Upper bar
Upper bar

Scroll bar Scroll bar


or chute bar
Scroll bar

Tributary mouth
bar or riffle
Lower bar
Lower bar
Lower bar

mst sst

Gamma ray Mean grain size

UPSTREAM PARTS OF CHANNEL BARS Upper bar

BANKFULL
channel Upper bar
Upper bar

Upper
Upper

bar
LEVEL

bar Unit bar


Cross-bar

Bar-head
unit bar

Unit bar
Unit bar Unit bar

Lower bar
Lower bar

Lower bar
Lower bar
Lower bar

Unit bar
mst sst
Gamma ray Mean grain size

UPSTREAM PART CHANNEL FILL CHANNEL FILL WITH DOWNSTREAM PART


OF CHANNEL FILL BANK SLUMPS OF CHANNEL FILL
BANKFULL LEVEL
Bank slumps
Unit bar

Unit bar
Lower bar

Lower bar

Lower bar
Lower bar
Unit bar

Gamma ray Mean grain size

FIG. 21.—Qualitative models of channel deposits. Vertical logs of typical sedimentary sequences through different parts of compound
bar deposits and channel fills for sandy rivers.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 107

whereas those associated with tributary-mouth bars occur nearer of accretion on progressively smaller bars (e.g., rows of unit
the base (Fig. 20). bars) as discharge is reduced (Figs. 21, 24). Small deltas may
Relatively small channels cutting into the upper parts of prograde into entrances of abandoned channels containing
channel bars, particularly between unit bars, are cross-bar chan- ponded water, thereby producing coarsening-upward se-
nels. Cross-bar channels commonly develop their own bars, the quences. Sediment gravity flows from cut banks may accumu-
geometry of which is controlled by the conditions of flow and late in the channel base as poorly sorted, structureless deposits
sediment transport in these channels. Where a cross-bar channel (Fig. 21). The suspended-load deposits drape over existing bed
joins a main channel, solitary delta-like deposits with avalanche topography. Horizontal suspended-load deposits lap onto in-
faces commonly form (e.g., chute bars; Fig. 23). The maximum clined channel margins (Fig. 20). In humid climates, peat may
depth of cross-bar channels is commonly less than a third of the accumulate in the ponded water of channel fills. In arid cli-
maximum depth of main channels, but any cross-bar channel mates, evaporitic tufas may form.
could be enlarged to become a main channel (e.g., chute cutoff). Channel-fill deposits grade laterally into channel-bar depos-
its, and it may be difficult to distinguish them in subsurface
Channel-Fill Deposits.— deposits. Channel-fill sequences can look very similar to channel-
bar-tail deposits. The deposits of the relatively small bars within
The deposits of channel fills are dependent on the history of channel fills may look similar to the deposits within cross-bar
flow through the channel following the beginning of abandon- channels. The fine-grained parts of channel fills may look very
ment. Channel abandonment is normally initiated by movement similar to overbank deposits, including lacustrine deposits.
of a bar into the entrance of the channel. If the angle between the
enlarging channel and the filling channel is relatively small, as in Evidence of Falling Flow Stage.—
low-sinuosity rivers, flow is maintained in the filling channel so
that bed load can be deposited (Fig. 17). Such bed load is com- Evidence of falling stage in channel deposits includes fining
monly moved as unit bars with superimposed ripples or dunes. of grain size and associated changes in sedimentary structures
Although bed load may extend a considerable way into such in the upper parts of large-scale inclined strata (Figs. 19, 21, 22).
filling channels, the downstream ends receive mainly fine-grained Cross strata associated with dunes and unit bars may have
suspended sediment and organic matter from slowly moving current ripples and wave ripples superimposed, and possibly
water. With larger angles of divergence, both ends of the aban- mud drapes with abundant plant debris. Rill marks oriented
doned channel are quickly blocked (Fig. 17), and most of the parallel to depositional slopes represent falling-stage drainage
channel fill is relatively fine grained and organic rich because of channels, and cross-stratified sand wedges represent the small
deposition from suspension in ponded water. deltas that form as these channels flow into standing water (Fig.
Channel fills generally fine upward, reflecting progressively 25). Desiccation cracks occur in emergent mud drapes, and
weaker flows during filling (Figs. 19, 21). They also generally rooted plants can colonize areas exposed at low flow stage. The
fine downchannel as water flow decelerates in that direction. level of these features in channel sequences gives an indication
The relatively coarse bed-load deposits at the upstream end of of the low-stage level. Burrowing and surface-browsing ani-
the channel fill tend to fine upwards, inasmuch as they repre- mals are most active following floods, and escape burrows may
sent progradation of bar-tail deposits into the channel entrance occur within the flood deposits.
(Fig. 21). Bed-load deposits in channel fills may show evidence
Paleocurrent Orientations.—
LEGEND FOR SEDIMENTOLOGICAL LOGS Paleocurrent orientations recorded in channel deposits de-
pend on: (1) the orientation of the bed forms and associated
Mudstone with minor sandstone lenses, sedimentary structures that vary with bed-form type, their
vegetation-rich layers, root casts, burrows, position in the channel, and with river stage; and (2) what part
and desiccation cracks. Degree of
bioturbation increases upwards. of the channel bar or fill is preserved (Allen, 1966). The mean
orientation of structures like pebble imbrication and various
Lenticular to wavy bedding scales of cross strata generally correspond with local water-flow
with asymmetrical ripples directions. However, near banks, these paleocurrent indicators
tend to be oblique to local flow direction, as are the dips of angle-
Small-scale cross strata (set thickness of-repose cross strata in scroll bars. Local paleo-flow directions
< 0.02 m) with asymmetrical ripples may be associated with deposition over a range of paleo-flow
stages and strengths, and are not necessarily parallel to the
Planar strata orientation of the high-stage channels. Furthermore, it is ex-
pected that paleo-flow directions from the downstream parts of
ancient channel bars are preferentially preserved. As a result,
1 Medium-scale cross strata the mean paleocurrent azimuth for any particular structure in a
(set thickness drawn to scale) channel deposit may not be parallel to the mean channel orien-
tation, and the range of azimuths probably differs from the
meters

range of local channel orientations. Thus, great care must be


Sandy gravel-stone including exercised in interpreting local channel orientations (and chan-
intraformational mud clasts (black) nel sinuosity) from paleocurrent data (e.g., Allen, 1966).
0
vf f m c vc Porosity and Permeability of Channel Deposits.—
mud sand gravel
Porosity and permeability of channel deposits vary spatially
FIG. 21 (continued).—Legend for models on opposite page. with variation in texture and internal structure, and such varia-
108 JOHN S. BRIDGE

A West
10 m
East
0 0.0
Vertical exaggeration = 2

Channel fill Unit bars


Time (ns)

Depth (m)
100 Side-bar deposit Channel fill Side-bar deposit 6.75

2 meters

B land surface

upper bar

lower bar

C FIG. 22.—A) Large-scale inclined strata in GPR profiles through


braided and meandering river deposits. Profiles are oriented
across-channel. Upper profile (Sagavanirktok River, Alaska,
U.S.A.) shows compound bar deposit with large-scale in-
clined strata dipping to west and passing into a channel fill.
Variable inclination of large-scale strata is associated with
deposition on unit bars. Lower in the profile, a confluence fill
is bounded on both sides by side (point) bars. The large-scale
strata within the side bars increase in inclination, and their
basal erosion surfaces become deeper, towards the confluence
scour. Basal erosion surfaces of bars are marked by white
arrows. B) Lower profile (South Esk River, Scotland) shows
point-bar deposits with large-scale inclined strata dipping to
left. Basal erosion surface of point-bar deposits (marked by
black arrows) is about 3 m below land surface. Discordance
in inclination of large-scale strata is marked by d. Upper-bar
deposits have more laterally continuous radar reflections
than lower-bar deposits. C) Trench showing lower-bar de-
posits (medium-scale trough cross-stratified sand) overlain
by upper-bar deposits (small-scale cross-stratified and bur-
rowed sand interbedded with dark layers of vegetation-rich
silt). The dark layers are low-flow deposits, and define the
upper parts of large-scale inclined strata

tion occurs over different scales of strataset. Information on the 3-


D variation in porosity and permeability over this range of scales
is lacking in general. In some channel-bar deposits, porosity and
permeability decrease upwards with mean grain size. The poros-
ity and permeability of channel-bar deposits is also expected to
decrease downstream because bar-tail deposits are likely to be
finer grained than bar-head and mid-bar deposits. Furthermore,
the finest deposits occur as low-flow drapes within large-scale
inclined strata, and in the uppermost large-scale inclined strata of
bar tails. Low-permeability strata are also expected in relatively
fine-grained channel fills that are concentrated near the margins
of channel belts.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 109

A
Bar-head unit bars

Bar-tail scroll bar

Cross-bar channel

B C

a
Unit bar b

FIG. 23.—A) View looking upstream of a Sagavanirktok bar-tail scroll bar with cross-bar channels that pass to the left into mouth bars
within an inner bank swale (about 2 m wide). Two bar-head unit bars are in the background. B) Unit bar with steep downstream
face on a point bar, Congaree River, South Carolina, U.S.A. Unit bar is about 0.5 m high. C) Trench through front of unit bar in
Part B showing medium-scale cross strata (a) formed by dunes migrating over the unit bar. Cross strata associated with sand
avalanching down the steep front of the unit bar are limited in extent (b).

Models of Superimposed Channel Bars, of preservation of the lower parts of the eroded bar increases with
Channel Fills, and Channel Belts the vertical deposition rate relative to the lateral migration rate of
the superposed bar, and the variability of channel scour depth
The spatial distribution of the deposits of individual channel and bar thickness. The relative importance of the ratio of deposi-
bars and fills could not be included easily in the models above tion rate to lateral migration rate of bars and the variability of
because it is very difficult to predict how individual channel channel scour depths (bar heights) in controlling the amount of
segments and bars migrate and become preserved in channel belts. preservation of truncated bars can be assessed using Equation 1.
It is necessary to develop models that predict details of the deposits In general, the variability of channel scour depths is the main
of several adjacent bars and channel fills within channel belts. control.
Vertical superposition of channel-bar and channel-fill depos- Object-based stochastic models have been used to distribute
its in single channel belts can occur by superposition of a cross- channel deposits within channel belts (e.g., Tyler et al., 1994;
bar channel on a main-channel bar and by migration of one main- Webb, 1994, 1995; Webb and Anderson, 1996; Deutsch and Wang,
channel bar over another (Fig. 26). In the latter case, the degree of 1996; Holden et al., 1998; Deutsch and Tran, 2002). The common
preservation of the overridden bar depends on the relative eleva- approach in these models is: (1) define shapes of channels; (2)
tions of the two superposed basal erosion surfaces. The likelihood position a series of channels randomly within an aggrading
110 JOHN S. BRIDGE

20 m
0 0.0
top of channel fill C

Depth (m)
Time (ns)

channel margin
100 5.0

basal erosion surface

Vertical exaggeration = 5
cut bank active channel bank

land surface

basal erosion surface

10 meters
land surface
1m mid-channel bar
small bar in channel
(35 ns
TWTT) channel

basal erosion surface

FIG. 24.—Channel-fill deposits in across-channel GPR profiles. Upper profile (Sagavanirktok River, Alaska) shows compound large-
scale inclined strata (boundaries marked by small arrows) formed by individual gravelly unit bars that filled the channel. Within
these compound large-scale strata are smaller-scale large-scale inclined strata formed by episodic migration of unit bars. Channel
fill is approximately 4 m thick, and vertical exaggeration of profile is 5. Lower profile (Calamus River, Nebraska, U.S.A.) has high-
amplitude reflections (medium-scale cross-stratified sand) overlain by low-amplitude reflections (bioturbated small-scale cross-
stratified sand interbedded with vegetation-rich sand). Channel fill contains deposits of small bars. Vertical exaggeration of
profile is 3.

channel belt; and (3) define sedimentary facies, porosity, and Vertical superposition of channel-bar and channel-fill depos-
permeability within the channels. None of these approaches (e.g., its can also result from superposition of distinct channel belts
Fig. 27) correctly represents the nature of channel deposits in without intervening floodplain deposits (Fig. 26). In cores and
channel belts, which are in fact composed predominantly of parts well logs, it may be very difficult to distinguish superimposed
of channel bars with relatively minor volumes of channel fills (see channel bars and fills in a single channel belt with superimposed
below). It is necessary to define shapes of objects properly (Fig. channel belts (Fig. 26). The ability to make this distinction hinges
27). The information shown in Figure 6 will assist in the scaling of on the ability to correctly interpret the different superimposed
objects. Numerical simulation of channel deposits within channel scales of strataset, as explained below.
belts is in its infancy.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 111

dune floodplain
crest vegetation

flow tracks and trails mud


rills cracks

roots
delta
mud draping
vel ripple marks
water le
burrows
channel

FIG. 25.—Falling-stage features at channel margins.

Qualitative Interpretation of Ancient Channel Deposits (3) knowledge of how channel-bar and channel-fill deposits
appear in variously oriented 2-D sections. Figure 28 shows de-
In order to make the best use of the new qualitative deposi- scriptions of some well-exposed Miocene fluvial deposits from
tional models to interpret ancient channel deposits, it is necessary the Siwaliks of northern Pakistan that have allowed detailed
to have: (1) detailed descriptions of large outcrops; (2) thorough interpretation of depositional environment (Willis, 1993a, 1993b).
understanding of the geometry, flow, and sedimentary pro- Figure 29 shows other examples of relatively simple qualitative
cesses, and modes of migration of modern channels and bars; and interpretations of ancient channel deposits.

FIG. 26.—Superimposed channel deposits in channel belts. Thick lines are basal erosion surfaces of compound channel bars, thin lines
are large-scale inclined strata, and arrows are idealized paleocurrent directions relative to outcrop plane (down is out of outcrop).
A) Downstream migration and climbing of one bar over another bar. B) Superposition of channel bars and fills of different depth
and width. C) Superposition of channel bars of similar size but different sinuosity and orientation. D) Superposition of different
channel belts. E) Superposition of channel bars in different channel belts and within channel belts. It may be difficult to distinguish
these two different types of superposition using vertical sedimentary logs or gamma-ray logs (compare two logs on right side of
figure). (A to D from Willis, 1993a; E from Bridge and Tye, 2000).
112 JOHN S. BRIDGE

A B
A B

C
C

FIG. 27.—Object models for channel belts from A) Tyler et al. (1994) and B) Deutsch and Wang (1996) Both representations are
unrealistic because real channel belts are composed of the deposits of channel bars and channel fills. C) More realistic
representation of objects (plans and cross sections of channel bars and adjacent channel fills) that must be distributed within
channel belts.

Quantitative Interpretation of value unless the grains were close to these thresholds at the time
Ancient Channel Deposits from Outcrops of deposition. Komar (1996) pointed out potential errors in some
of the methods for determining threshold bed shear stress for
The most sophisticated quantitative interpretation of channel movement of gravel. Grain-size data have yet to be used seriously
geometry, flow, sediment transport, and migration requires the to quantitatively estimate flow parameters such as bed shear
same kind of information mentioned above for qualitative inter- stress during transport and deposition. Sedimentary structures
pretation, but it must be quantitative. The quantitative models and preserved bed forms indicate the geometry of bed forms that
described above have been used to interpret ancient meandering- existed at the time of deposition. If these bed forms are assumed
river deposits from the Devonian of southern Ireland and New to have been in equilibrium with the flow, estimates can be made
York State (Figure 30; e.g., Bridge and Diemer, 1983; Bridge and of ranges of bed shear stress or flow velocity and depth. The mean
Gordon, 1985; Gordon and Bridge, 1987; Willis, 1993c) and an- thickness of various scales of cross strata can give estimates of the
cient braided rivers in the Miocene Siwaliks of northern Pakistan mean heights of the bed forms (e.g., ripples, dunes, bars) respon-
(Willis, 1993a, 1993b; Khan et al., 1997; Zaleha, 1997b, 1997c). sible, and the heights of dunes and bars can be related to flow
Willis (1993a, 1993b) was able to quantitatively reconstruct the depth. Estimates of flow velocity, depth, and friction coefficients
width, depth, mean velocity, slope, wavelength, and sinuosity of from grain-size data and reconstructed bed forms have been used
individual channel segments in these Siwalik deposits, and, to estimate channel slope using formulae for uniform flow.
because of the excellent exposures, to estimate channel-belt widths Unfortunately, the flow equations used for these procedures are
and degree of braiding. Channel bars were interpreted to have commonly misapplied or are inappropriate. It is also commonly
migrated mainly by downstream translation and bend expan- difficult to understand what the reconstructed flow velocities
sion, but also by channel switching within the channel belts. and depths actually mean. Are they local or spatially averaged
Other, less sophisticated methods of quantitative interpreta- values, and what flow stages do they represent?
tion of paleochannel hydraulics and geometry are routinely Average sinuosity of paleochannels can be estimated from the
applied (reviews in Bridge, 1978; Ethridge and Schumm, 1978; maximum range of paleocurrent directions observed in a single
Williams, 1988; North, 1996). Grain-size data have yielded esti- channel-belt deposit, provided that paleocurrent indicators are
mates of the threshold bed shear stress for bed-load or sus- analyzed carefully. Paleocurrent indicators should represent lo-
pended-load motion, although this information is of limited cal paleochannel direction (e.g., medium-scale trough cross strata
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 113

FIG. 28.—Example of Miocene Siwalik deposits of northern Pakistan. Position of photo is shown as box in figure to right, which shows
channel sandstone bodies (stippled) with large-scale inclined strata, floodplain mudstones (unshaded), and well-developed
paleosols (vertical line ornament). Lower figure gives more sedimentological details of the extensive sandstone body shown in
the upper right figure (which includes logs 5 to 18). Line diagrams modified from Willis (1993a).

from lower-bar deposits), and the preferential preservation of fraught with problems, including inadequate empirical regres-
bar-tail deposits must be recognized. Calculation of sinuosity sion equations and their misapplication (discussed in Bridge,
from paleocurrent ranges requires a functional relationship be- 2003). Unfortunately, such empirical equations that relate geo-
tween these two parameters. metrical and sedimentary characteristics of ancient rivers to
On Holocene to Pleistocene floodplains, it is commonly their paleodischarge have largely formed the basis for the field
possible to observe the geometry and plan form of paleo- of paleohydrology. The hydrology of major paleofloods has
channels. In some well-exposed fluvial deposits, the width and gained attention recently (e.g., Baker et al., 1988; Martini et al.,
depth of channels and bars, and (exceptionally) the length and 2002). One novel way of assessing paleoflood levels is to deter-
sinuosity of channel bends, can be observed directly (reviews in mine the level of backwater deposits in canyons (Kochel and
Bridge, 1978; North, 1996). In most cases, only one or two of Baker, 1988). The depth and velocity of floods associated with
these geometric parameters can be observed, and empirical catastrophic draining of ice-dammed lakes have been inter-
equations derived from modern rivers have been used to pre- preted by examining large bedforms on floodplains (e.g., chan-
dict other geometrical parameters. Observed or calculated geo- nelized scablands; Baker and Nummedal, 1978).
metric parameters of rivers are then used, in some cases along
with sedimentary data, to calculate channel-forming discharge Estimation of Paleochannel Depth from Subsurface Deposits
using empirical regression equations (reviews in Dury, 1976,
1985; Bridge, 1978; Ethridge and Schumm, 1978; Williams, 1988; It is desirable to estimate the thickness and width of subsur-
Bridge and Mackey, 1993b; North, 1996). This procedure is face channel-belt deposits in view of their importance as aqui-
114 JOHN S. BRIDGE

FIG. 29.—A) Devonian river channel deposits from SW Ireland with large-scale inclined strata dipping to left (top of cliff) and fine-
grained channel fills. Person in lower left for scale. B) Carboniferous river channel deposits from northwest Germany, showing
large-scale inclined strata dipping to left. Basal erosion surface is immediately above head of person.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 115

Plan view of channel centerlines Plan view of channel centerlines

FIG. 30.—Quantitative interpretation of Devonian channel-bar deposits from New York State, U.S.A. Photo on left shows a channel-
bar deposit viewed parallel to flow direction (basal erosion surface marked by arrow). Reconstructed channel-bar geometry and
migration pattern are show below. Photo and line diagrams to right show a channel bar viewed approximately normal to flow
direction (arrow indicates basal erosion surface). Channel reconstructions from Willis (1993c).

fers and hydrocarbon reservoirs (Bridge and Tye, 2000). The of the range of maximum channel depths. An independent
first step is estimation of maximum paleochannel depth. In check on the estimates of bank-full flow depth is to use the
order to do this, channel-belt sands and gravels must be distin- relationships between thickness of medium-scale cross sets,
guished from floodplain sands, and the various scales of strataset dunes and flow depth (Fig. 31; Bridge, 2003).
must be distinguished, particularly those due to individual
floods, single channel bars and fills within one channel belt, and Estimation of Widths of Single Channel Belts
individual channel belts (method shown in Figure 31). This from Subsurface Deposits
requires knowledge of spatial variations in grain size, sedimen-
tary structures, paleocurrents, and degree of stratal disruption Four commonly used methods for estimating the geometry of
(determined from cores and well logs). However, it is difficult isolated channel belts are: (1) well-to-well correlation; (2) using
to distinguish meters-thick fining-upward sequences associ- empirical equations relating maximum channel depth, channel
ated with sandy to muddy fills of main channels from a rela- width, and channel-belt width; (3) measurement of outcrop ana-
tively thin sequence of overbank sandstone (e.g., levee, crevasse logs; and (4) amplitude analysis of 3-D seismic horizon slices.
splay, lacustrine delta) overlain by muddy floodbasin deposits. These are discussed in turn.
Also, upper-bar deposits look very similar to near-channel
floodplain deposits. Superimposed channel bars or fills are Well-to-Well Correlation.—
difficult to distinguish from single-channel bars or fills. The
thickness of all single, nontruncated channel bars or fills (from Well-to-well correlation of channel-belt sandstone bodies
the tops of channel belts) must be measured to get an estimate using wireline logs has been the most common method for
116 JOHN S. BRIDGE

and ancient channel belts have width/maximum channel depth


of between 700 and 20 (Bridge and Mackey, 1993b). It is com-
monly stated that this ratio is larger for braided rivers than for
meandering rivers. This is a moot point when utilizing core and
wireline log data, because such a distinction between
paleochannel patterns cannot be made. Furthermore, this sup-
position is not generally correct. For example, the channel-belt
width/maximum bankfull depth for the meandering lower
Mississippi River approximates that of the braided Brahmaputra
River, being about 300.

Empirical Equations.—
In order to make the critical assessment of expected width
of channel-belt sandstone bodies, it is first necessary to deter-
mine whether the sandstone body is a single channel belt or a
combination of channel belts. The widths of single channel
belts can be estimated using empirical equations that relate
maximum channel depth, channel width, and channel-belt
width (Bridge and Mackey, 1993b; Bridge and Tye, 2000). This
approach requires estimates of maximum bankfull channel
depth from cores or well logs. Because the empirical equations
available have large standard errors, estimates of channel-belt
width are imprecise.

Outcrop Analogs.—

The use of outcrop analogs to interpret subsurface strata is


very popular but has serious shortcomings. Obviously, the inter-
preted depositional environments of the outcrop analog and the
subsurface strata must match. It is difficult to make detailed
interpretations of depositional environments from typical sub-
surface data, and outcrop data can easily be misinterpreted.
Outcrops are rarely extensive or numerous enough to allow
unambiguous determination of the three-dimensional geometry
FIG. 31.—Estimation of flow depth from cores and gamma-ray and orientation of channels and channel belts. This is why it is
logs through channel-bar deposits (from Bridge and Tye, desirable to use analog data from Holocene depositional environ-
2000). Maximum bankfull flow depth can be estimated from ments, where channel-belt dimensions can be determined easily,
the thickness of channel-bar deposits (7.5 m). This estimate and the relationship between the nature of the deposits and the
can be checked using the mean thickness of medium-scale geometry, flow, and sedimentary processes of the environment
cross sets (formed by dunes) from the lower part of bar can be established unambiguously.
deposits (0.29 m), giving mean dune height of about 0.85 m.
This dune height is appropriate for estimated water depths Amplitude Analysis of 3-D Seismic Horizon Slices
during formation of 6 to 7 m.
Amplitude analysis of 3-D seismic horizon slices is the only
method capable of yielding directly the width of channel belts,
estimating channel-belt widths and orientations (Fig. 32). The and imaging the channel pattern (sinuosity, channel splitting)
spatial resolution of this technique can be no better than the of subsurface sandstone bodies (Fig. 33). This is also the only
average well spacing. The validity of this technique is very method that can be used to predict the spatial distribution of
much dependent on the correlation rules utilized. Once a suit- channel-belt thickness and lithofacies. These are major ad-
able horizontal datum has been chosen for the wells to be vances. However, this method depends on the resolution of the
correlated, it is necessary to establish whether sandstone bodies seismic data relative to the thickness of the sandstone bodies
at similar stratigraphic levels in different wells can be corre- imaged, and requires calibration by wireline logs and cores. In
lated. In order to make this assessment, it is essential to have a general, sandstone-body thickness must be greater than ap-
reasonable genetic interpretation of the sandstone body, and a proximately 10 m.
model for its lateral extent and lateral variation in thickness and
lithofacies. Well-to-well correlation is commonly compromised Estimation of Width of Superimposed Channel Belts
by lack of a realistic model for the possible lateral extent and from Subsurface Deposits
lateral variation of sandstone bodies, and erroneous assump-
tions such as: (1) sandstone bodies positioned at the same Widths of superimposed channel belts can be estimated with
stratigraphic level must be connected between adjacent wells; the help of alluvial stratigraphy models (Bridge and Mackey,
(2) vertical sequences through channel deposits indicate the 1993b; Mackey and Bridge, 1995), and depend on the proportion
paleochannel pattern and hence the geometry of channel-belt and degree of connectedness of channel-belt deposits in a cross
sandstone bodies; and (3) ratios of sandstone-body width to section (Fig. 34). For channel-deposit proportion less than about
thickness are closely related to paleochannel pattern. Modern 0.4, channel belts are unconnected and sandstone-body width
Permian
salt

coal

Shale

coal

coal
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS

coal
coal
Sandstone

FIG. 32.—Channel-belt width is commonly estimated using well-to-well correlation of channel-belt deposits if channel belts are wider than well spacing. This requires
an estimate of expected channel-belt width from channel-bar thickness. However, it is difficult to distinguish relatively thick floodplain sands from thin channel-bar
deposits and the sandy parts of channel fills. Fine-grained channel-fill deposits are difficult to distinguish from floodplain shale. In this figure, the deposits are
interpreted as channel-belt deposits (stippled), floodplain sands (stippled), or floodbasin shales (unstippled). The datum used to aid correlation is the coal seam in
the middle of the section. Channel-belt deposits (letters A to P) were recognized on the basis of their (large) thickness and gamma-ray (GR) patterns. Correlation of
channel deposits was based on estimated width derived from channel-bar thickness, hence maximum channel depth. To do this it is not necessary to know whether
the channel was meandering or braided. Also, channel fills and floodplain sands help define channel-belt edges (e.g., L and H). Superimposed channel bars are
distinguished using GR patterns (e.g., J). Floodplain sandstones are recognized and correlated on the basis of their (small) thickness, expected geometry, and
relationship to channel-belt edges.
117
118 JOHN S. BRIDGE

FIG. 33.—Amplitude analysis of a 3-D seismic horizon slice showing the width and channel pattern of channel-belt sandstone bodies.
Cross section (right) shows correlated logs and position of horizon slice. Log 3 cuts the variable-width, straight channel belt (red
in left figure). Logs 4 to 6 cut through a point bar and channel fill of a slightly older channel belt (green and blue in left figure).

CDP = channel-belt proportion


Vertical exaggeration ~ 10

FIG. 34.—Channel-belt connectedness increases with channel-deposit proportion (CDP). For CDP < 0.4, most channel belts (shown
as stippled boxes in the cross-floodplain section to the upper right) are unconnected, such that frequency distributions (to upper
left) of channel-deposit width or thickness (relative to floodplain width or thickness) are bimodal with a large mode equivalent
to unconnected channel belts. As CDP increases, more channel belts are connected, channel deposits become larger, and the
frequency distributions of channel-deposit width or thickness becomes polymodal. For CDP > 0.75 (lower figures), all channel
belts are connected and the single-channel deposit is as wide and thick as the floodplain.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 119

equals channel-belt width. As channel-deposit proportion in- induces deposition by decelerating sediment-laden flows, and
creases, some channel belts become connected, the mean and protects surface sediment from entrainment by wind or water.
standard deviation of width increase, and the frequency distribu- Plant cover is sparse in arid and semiarid climates, such that
tion becomes polymodal. If channel-deposit proportion exceeds surface sediment may be moved by the wind. The geometry of
about 0.75, all channel belts are connected, and the single sand- floodplains changes in time and space as a result of: channel
stone body has a width equal to floodplain width. migration within the channel belt; migration, cutting, and filling
of floodplain channels; large-scale movements of channel belts
FLOODPLAINS (avulsions); local tectonism; and progressive deposition or ero-
sion.
A floodplain is a strip of land that borders a stream channel The fastest flow in simple channel–floodplain systems is in
and that is normally inundated during seasonal floods. Flood- the main channels, and flow velocity and bed shear stress on the
plains develop in all alluvial valleys and on alluvial fans and floodplain diminish away from channel margins (Fig. 36). The
deltas, irrespective of the channel pattern. The interaction be- most abrupt lateral decrease in flow velocity occurs at the chan-
tween water flow, sediment transport, and bed topography of nel–floodplain margin, related to vortices with horizontal axes
floodplains is not known as well as for channels (review by spaced periodically along the channel margins that cause local
Bridge, 2003). Comprehensive field studies of water flow and exchange of water between the channel and the floodplain. As the
sediment transport over floodplains during overbank floods do depth of flow on the floodplain increases relative to that in the
not exist, mainly because of difficulties of observation and lack of main channel, the flow velocity on the floodplain also increases,
students willing to make the observations (but see Velikanov and and the reduction in velocity at the channel margin is not as great
Yarnykh, 1970; Hughes and Lewin, 1982; Nicholas and McLelland, as with shallow floodplain flows (Fig. 36). With wide floodplains
1999). Water flow and sediment transport on floodplains is (width/depth > 100), the velocity of the floodplain flow decreases
complicated by variable floodplain width and surface topogra- with distance from the channel only in a zone close to the channel
phy (channels, depressions, mounds of sediment such as levees margin (Wormleaton, 1996). Also, flow velocities on floodplains
and crevasse splays, vegetation, and structures produced by are relatively high in zones of flow convergence and relatively
humans and other animals). Many experimental studies of low in expanding flow zones, and are greatly influenced by
overbank flow adjacent to river channels have been undertaken variations in bed roughness.
(recent examples are Knight and Shiono, 1996; Naish and Sellin, Sediment is transported over the floodplain as bed load and
1996; Sellin and Willetts, 1996; Willetts and Rameshwaran, 1996; suspended load during floods. The sediment comes from the
Wormleaton, 1996; Lai et al., 2000; Patra and Kar, 2000; Knight main channel and tributary channels, the valley sides, and the
and Brown, 2001; Myers et al., 2001; Valentine et al., 2001), but floodplain itself. A large range of sediment size is commonly
mostly with steady flows over simple channel–floodplain geom- available, from mud to gravel. The coarser-grained sediment may
etry, and with immobile boundaries without sediment move- be in the form of mud pellets (consolidated by desiccation or
ment. Numerical models of floodplain flow and sediment trans- bioturbation), soil concretions, and organic debris (bones, shells,
port are inadequate at present, and there are no quantitative, 3-D plant axes). Indeed, much of the mud that is transported on
depositional models for floodplains. Qualitative models of flood- floodplains may be in the form of pellets. Sediment is routed onto
plain deposits are also relatively poorly developed, as demon- the floodplain from the main and tributary channels via smaller
strated below. channels, sheet flows, and the large-scale vortices at channel
margins. Bedload sand is transported mainly as ripples and
General Patterns of Geometry, Water Flow, upper-stage plane beds. However, dunes occur also, especially in
Sediment Transport, Erosion, and Deposition floodplain channels, and antidunes occur in very rapid, shallow
flows.
Floodplain geometry has been reviewed by Allen (1965, 1970), Changes in water flow and sediment transport over flood
Brierley et al. (1997), and Bridge (2003). Floodplains normally periods control erosion and deposition on floodplains (review in
contain active and abandoned alluvial ridges that rise several Bridge, 2003). In general, water initially gets to the floodplain
decimeters to meters above adjacent lowlands (flood basins). Allu- during rising flow stage through crevasse channels, through low
vial ridges contain active and abandoned channels and bars (the parts of levees, and by overland flow from the valley margins. At
channel belt), levees, and crevasse channels and splays (Fig. 35). this stage, the water level in the main channels may be much
Levees are discontinuous, wedge-shaped ridges around active higher than that in the flood basin, leading to accelerating, erosive
and abandoned channels (Fig. 35). Levees commonly have chan- flows from the channel to the flood basin. During peak flood, water
nels cut into their surfaces. The larger ones are called crevasse completely covers the floodplain and flows more or less down
channels and split downslope into smaller distributaries sur- valley. The water flows in a very broad, shallow channel (the
mounting fan- or lobe-shaped mounds of sediment called crevasse floodplain) in which there are smaller, deeper channels (the main
splays (Fig. 35). In some cases, levees comprise a series of adjacent channels and crevasse channels). The flow patterns on the flood-
crevasse splays. Crevasse channels can have their own levees and plain during this stage are described above and in Figure 36.
terminal mouth bars. The distal margins of crevasse splays can During falling flow stage, water and sediment flow back into the
either thin gradually or end abruptly with a steep (angle-of- main channels through floodplain drainage channels, and as
repose) slope. Crevasse splays that terminate in permanent lakes ground water. Floodplain lakes gradually diminish as water level
are similar to lacustrine deltas. Flood basins are commonly seg- goes down. Freshly exposed sediment surfaces are modified by
mented into subsidiary basins by crevasse splays, alluvial ridges the wind, by plant growth, and by the activities of animals.
of tributary channels, or abandoned alluvial ridges. Permanent Desiccation cracks appear in muddy sediments, and in arid
lakes and marshes may be present in wet climates, whereas lakes climates salts may be water precipitated in the soil as a result of
are ephemeral in dry climates. Permanent lakes are particularly evaporation of surface and ground water.
common on coastal plains, and where there is local tectonic The greatest floodplain deposition rates of the coarsest sedi-
subsidence or base-level rise. Floodplain drainage channels are ments generally occur near the margins of channels (on levees
common, and those that flow into lakes form deltas. Plant cover and crevasse splays, thus explaining the origin of alluvial ridges)
120 JOHN S. BRIDGE

FIG. 35.—Geometry of floodplains. A) Alluvial ridge with active channel and levee (background) and abandoned channels. Paraná
River, Argentina. Photo from J. Best. B) Active channel and levee (left) passing to right into wet floodbasin. Cumberland Marshes,
Saskatchewan, Canada. Photo from N. Smith. C) Active crevasse splay from Saskatchewan River, Canada. Floodbasin is bordered
by older channel belt (background). Cumberland Marshes. Photo from N. Smith. D) Crevasse channel and crevasse splay passing
into dry floodbasin (background). Brahmaputra River. E) Desiccated muddy floodbasin with channel belt in background.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 121

flow velocity vector

floodplain channel
1
flow deep
velocity
shallow
0
FIG. 36.—Idealized overbank flow patterns on floodplains (based on information in Knight and Shiono, 1996; Wormleaton, 1996).
Perspective diagram indicates decrease in downvalley flow velocity with distance away from channel, and large-scale vortices
at the channel margin. Across-floodplain variation in flow velocity depends on the depth of overbank flow relative to channel
depth. The interaction between overbank flow and channel flow pattern is not shown.

because of rapid spatial deceleration and decreasing turbulence main channel are similar to upper-bar deposits but decrease in
intensity (see references in Bridge, 2003, p. 268). Relatively high grain size (and change internal structure) with distance from
rates of floodplain deposition also occur in zones of flow decel- channels. Layers of drifted vegetation are common in the overbank
eration such as abandoned channels and floodplain lakes. The deposits of humid climates, and they are generally bioturbated by
lowest deposition rates of the finest sediments occur in flood animals and plant roots. The stratasets may be sheet-like, wedge-
basins distant from channels and areas that are relatively el- shaped, or lenticular, depending on the local environment of
evated. Although the grain size of sand on modern floodplain deposition.
surfaces tends to decrease away from the channel belt, mud
usually accumulates as a more or less continuous blanket (ref- Qualitative Depositional Models for Floodplain Deposits
erences in Bridge, 2003, p. 269). Deposition rate (averaged over
hundreds of years) on modern floodplains decreases exponen- Levees, Crevasse Splays, and Lacustrine Deltas.—
tially with distance from the active channel belt. This means that
the cross-valley floodplain slope and elevation of alluvial ridges Depositional models for levees, crevasse splays, lacustrine
above flood basins increase with time. The average thickness of deltas, and floodbasins are still rudimentary, because of the lack
sediment deposited on floodplains during seasonal floods is on of detailed study (Fig. 38; review by Bridge, 2003). However,
the order of millimeters to centimeters. However, floodplain some recent studies have added greatly to our knowledge, nota-
deposition rate averaged over millions of years is on the order bly work on crevasse splays and lacustrine deltas in the
of 0.1 mm per year, because long-term floodplain deposition is Cumberland Marshes in Saskatchewan, Canada (Fig. 39; Smith et
interrupted by lateral shifts of alluvial ridges (avulsions) and by al., 1989; Smith et al., 1998; Smith and Perez-Arlucea, 1994; Perez-
periods of erosion (Bridge and Leeder, 1979; Sadler, 1981; Enos, Arlucea and Smith, 1999; Morozova and Smith, 1999, 2000; Farrell,
1991). Erosion occurs where flow is accelerated in locally nar- 2001), on the lacustrine deltas of the lower Mississippi Valley (Tye
row or topographically high floodplain sections and where and Coleman, 1989a, 1989b), and on a crevasse splay in Nebraska
vegetation cover is poor. (Fig. 39; Bristow et al., 1999). However, there is still a need for
The basic sedimentation units on floodplains are millimeter- more detailed descriptions of the 3-D variation of strataset geom-
to decimeter-thick stratasets deposited during floods (Figs. 7, 37, etry, grain size, sedimentary structures, paleocurrents, and or-
38; references in Bridge, 2003, p. 270). Basal erosion surfaces are ganic remains.
present if erosion preceded deposition. Upward-fining stratasets The 3-D geometry of crevasse-splay deposits is different from
indicate deposition during temporally decelerating flows, whereas that of levee deposits (Fig. 38). Crevasse-splay deposits tend to be
those that coarsen upwards then fine upwards reflect deposition coarser grained and thicker than levee deposits, although the
during accelerating then decelerating flows. Grain sizes and edges of crevasse-splay deposits farthest from the main channel
internal structures depend on local flow conditions and sediment may be difficult to distinguish from levee deposits. Flood-gener-
availability. Many floodplain deposits are planar-stratified and ated stratasets of crevasse splays are similar to those of levees, but
small-scale cross-stratified fine to very fine sands interbedded medium-scale cross strata (formed by dune migration) are more
with silt and clay. In general, overbank deposits closest to the common in sandy crevasse-splay deposits. Channel-bar and chan-
122 JOHN S. BRIDGE

FIG. 37.—Typical flood-generated sedimentation unit on a floodplain near the main channel (Brahmaputra River, Bangladesh). Planar
laminae overlie an erosion surface (not shown), in turn overlain by small-scale cross laminae (climbing-ripple type). Load
structures are common in the small-scale cross sets. Sedimentation unit fines upward and is capped by bioturbated mud.

nel-fill deposits are common in crevasse splays (Fig. 38), and Channel Fills.—
these may be difficult to distinguish from the main-channel
deposits. The channels on crevasse splays are expected to be Channel fills in floodplain deposits may be associated with
smaller on average than those in main channels, and show abandoned main channels, crevasse channels (including flood-
evidence of periodic cessation of discharge (e.g., desiccation- plain drainage channels), and tributary channels. After their
cracked mud layers, root casts, burrows, tracks and trails through- ends become plugged with bed-load sediment, these aban-
out). However, some crevasse channels may be similar in size to doned channels become lakes, and receive mainly suspended-
main channels (especially immediately prior to an avulsion), and load sediment. Deposits are typically small-scale cross-strati-
some main channels may be ephemeral just like crevasse chan- fied sands and muds. Lacustrine deltas may form at the en-
nels. The margins of levees, crevasse splays, and lacustrine deltas trances to these lakes. Channel fills normally contain abundant
can slope at up to the angle of repose. If they reach the angle of plant debris, shells of freshwater molluscs, vertebrate bones
repose, the marginal deposits resemble those of Gilbert-type (given the appropriate climate and stage in earth history),
deltas (Figs. 39, 41). burrows, tracks, and trails (see below).
Within levees, crevasse splays, and lacustrine deltas, groups of
flood-generated stratasets may occur in distinctive vertical se- Flood Basins.—
quences that are up to meters thick and perhaps hundreds of
meters in lateral extent (Figs. 38, 39, 41). Coarsening-upward Flood-basin deposits are the finest sediments available for
sequences are produced by progradation of the sediment bodies transport (e.g., silts and clays), with subordinate millimeter- to
into flood basins or lakes, and upward-fining sequences are pro- centimeter-thick sheets and lenses of sand containing small-
duced by abandonment. Such progradation and abandonment scale cross strata and both wave-ripple and current-ripple marks
may be associated with migration and abandonment of individual (Fig. 40). Evidence of subaerial exposure in flood-basin deposits
channels in a crevasse splay or lacustrine delta, migration and is desiccation cracks and raindrop imprints in mud, and wind-
cutoff of channels within the active channel belt, or avulsion of the blown sand. Burrows and root casts, and layers of drifted plant
whole channel belt. However, these sequences might also be material, may be abundant depending upon climate. However,
related to regional changes in sediment supply and deposition rate exceptional sheet floods may deposit thick (up to a meter or so)
associated with, for example, climate change, tectonism, or relative sheets of sediment over large areas of a floodplain. Internal
sea-level changes. Distinguishing among these various possibili- structures may be medium-scale cross strata and planar strata,
ties is no trivial task, requiring observation of floodplain deposits formed from dune migration and upper-stage plane beds, re-
of a given age across the full extent of the floodplain. spectively.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 123

B C

Channel Levee

Levee

D E

FIG. 38.—A) Simple depositional models for levees and crevasse splays (from Bridge, 2003). Levee deposits from B) Devonian of New
York, U.S.A., and C) Miocene of northern Pakistan. Levee deposits thin away from main channels. D, E ) Crevasse-splay deposits
from the Carboniferous of eastern Kentucky, U.S.A. Sections are normal to flow direction, and show channels cutting through
lobate sandstone bodies (a).

Floodplain Lakes.— Common shelly fossils are pelecypods, gastropods, and ostra-
cods. If sediment supply is low, chemical or biochemical pre-
Stratasets formed in lakes are sheet-like and millimeters to cipitation of deposits may be important. For example, carbon-
centimeters thick. Evidence for waves in lakes includes cross- ate mud may be formed by calcareous cyanobacteria. A com-
stratified sands and silts with relatively small wave-ripple mon association of features in such deposits is centimeter-
marks, and planar laminae (Fig. 40). Burrows, tracks, and trails thick strata of calcite or dolomite mud with pellets, ostracods,
are common as long as the lakes are oxygenated and not burrows, and evidence of cyanobacterial filaments and mats.
hypersaline, and root traces occur in shallow-water deposits. In swampy areas with low sediment supply in temperate and
124 JOHN S. BRIDGE

SOUTH NORTH
standing
dunes in exposed areas ripples water
mud deposition
splay channel

flow direction
dunes in distal slip face
crevasse channel vegetated levee

SOUTH NORTH
crevasse channels prograding splay deposits

backswamp deposits

WEST EAST
mud deposition crevasse channel

2m
backswamp deposits local channel incision
20m

LEGEND
medium-scale trough cross
pre-splay deposits strata (formed by dunes) trees
curved-crested dunes fish in water
mud (silt and clay) current ripples and associated
shrubs and grass
small-scale cross strata
sand cross strata on distal slip faces roots

FIG. 39.—Depositional models for crevasse splays from Bristow et al. (1999) based on the Niobrara River, Nebraska, U.S.A. (this page),
and from Perez-Arlucea and Smith (1999) based on Cumberland Marshes, Canada (opposite page).

humid climates, peat may accumulate. In arid climates, evapor- ment supply as well as depth and areal extent of the lake (e.g.,
ites may form. vertical changes in grain size, composition, internal structure,
Upward-coarsening sequences (decimeters to meters thick, and thickness of strata). As with other overbank deposits, such
and up to kilometers across) are expected from progradation of changes may be local and associated with migration of channels,
lacustrine shorelines, whereas fining-upward sequences are ex- levees, and crevasse splays, local tectonism, or regional and
pected from retrogradation (e.g., Fig. 41). Lacustrine deposits associated with changes in climate, relative sea level, and tec-
also commonly show evidence for temporal changes in the sedi- tonism.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 125

FIG. 39 (continued).—
126 JOHN S. BRIDGE

FIG. 40.—Floodbasin deposits. Sandstone-filled desiccation cracks in mud (now eroded) that was draping small wave-ripple marks
(wavelengths on the order of a centimeter) and current-ripple marks. The wave-ripple marks are symmetrical at top left but
become asymmetrical and smaller in wavelength towards the lower right. Their crest lines also change orientation as they become
more asymmetrical. These changes in wave-ripple marks are associated with shoaling water at the edge of an ephemeral
floodplain pond. Current ripples in lower half of photo indicate a unidirectional flow to the right.

Qualitative Models for Floodplain Soils comprises the lateral variants of a soil. A common type of catena
on floodplains is associated with a decrease in elevation, deposi-
Alluvial soil features were recently reviewed by Retallack tion rate, and grain size with distance from the channel belt (Fig.
(1997, 2001), Kraus (1999), and Bridge (2003). Typical features of 44; e.g., Fastovsky and McSweeney, 1987; Platt and Keller, 1992;
alluvial soils include: (1) horizons (e.g., A, B, C); (2) textural Wright and Platt, 1995; Zaleha, 1997a; Cojan, 1999; Kraus and
evidence of leaching of soluble materials and eluviation of clays Aslan, 1999; Wright, 1999). Soils on sandy, well-drained levees
from A horizons and precipitation of secondary minerals and and crevasse splays have a relatively thick oxidized and leached
illuviation of clays in B horizons; (3) disruption of original struc- zone (zone of aeration) underlain by a gleyed horizon (saturated
tures by burrowing organisms, plant roots, changes in moisture zone). The lower, muddy, poorly drained flood-basin deposits
content, and growth of secondary minerals; and (4) characteristic are more gleyed. If soils undergo extremes of wetting and drying,
coloration and mottling associated with chemical alteration of calcium carbonate is leached from the zone of aeration and
parent material and formation of new minerals (Fig. 42). The accumulates as glaebules in the capillary fringe above the water
degree of development of these features in soils depends on time, table. Poorly drained flood-basin soils may have glaebules of
deposition rate, climate, vegetation, topography relative to the both calcium carbonate and iron oxide. Local or widespread
water table, and source materials. Well-developed soils with hori- aggradation on floodplains may result in “drying-out” vertical
zons require on the order of 103 years to form, and relatively low sequences of soils (Fig. 45). Decreasing deposition rate from
deposition rate (less than the order of millimeters per year; Leeder, channel belt to floodbasin may also result in increasing degree of
1975). Such soils are typically decimeters to meters thick. Different soil development farther away from channel belts, and change in
types of soils and paleosols defined based on distinctive features degree of soil development in vertical sequences of paleosols has
are called pedofacies (e.g., Table 2 and Figure 43), although defini- been related to varying proximity of channel belts (Fig. 44; Bown
tions of modern pedofacies cannot always be applied to ancient and Kraus, 1987; Kraus, 1987; Kraus and Aslan, 1999). Further-
pedofacies, because of diagenesis (Mack et al., 1993). more, the relationship between increasing soil maturity and
Soils and paleosoils commonly vary laterally and in time, and decreasing deposition rate has been related to long-term, large-
various models have been proposed for such variations. A catena scale changes in deposition rate, as seen in a later section on
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 127

FIG. 41.—Lacustrine deposits. A) Coarsening-upward


sequence (about 4 m thick) from laminated shale
(a) to cross-laminated and ripple-marked sand-
stone and siltstone (b) to channel-fill sandstone
(c). From Carboniferous of Eastern Kentucky,
U.S.A. Interpreted as progradation of crevasse
splay into lake, analogous to Cumberland
Marshes (Fig. 39). B) Coarsening-upward se-
quence (about 4 m thick) from laminated shale (a)
to sandstone with angle-of-repose cross strata
(b). From the Carboniferous of northwestern
Germany. Interpreted as progradation of distal
slipface of a crevasse splay into a lake, analogous
to Niobrara example in Fig. 39.

B
128 JOHN S. BRIDGE

FIG. 42.—Paleosol features. A) nodular calcareous B horizon (a) overlain by noncalcareous mottled A horizon (b), overlain by stratified
non-pedogenic deposits (c). Soil profile is 2 to 3 meters thick. B) interconnected calcareous glaebules in a B horizon. Scale is 0.1
m long. Photos from Siwaliks of Pakistan, courtesy of Mike Zaleha.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 129

TABLE 2.—Classification of soils.

Entisol — Very weakly developed soils.


Inceptisol — Weakly developed soils.
Andisol — like Inceptisol but formed on volcanic ash.
Vertisol — Abundant swelling clay (mainly smectite) subjected to
seasonal extremes of wetting and drying (expansion and
shrinkage), commonly have blocky to columnar ped struc-
tures, cutans, slickensides, pseudo-anticlines, as well as cal-
careous and iron oxide glaebules.
Aridisol — Light color, calcareous layer close to surface (within a
meter) or evidence of precipitation of evaporite minerals such
as gypsum. May contain silcrete horizon. Desert soil.
Mollisol — Organic-rich A horizon with abundant roots and
burrows, and B horizon enriched in clay or calcium carbonate.
Grassland soil.
Histosol — O horizon (peat) with uncompacted thickness of at
least 0.4 m. Swamp soil.
Spodosol — Thick, well-differentiated horizons with B horizon
enriched in sesquioxides and organic matter, and little or no
clay or calcium carbonate. Forest and woodland soil.
Alfisol — Thick, well-differentiated horizons with B horizon
enriched in clay, red sesquioxides, and cations of Ca, Mg, Na,
K. Forest and woodland soil.
Ultisol — similar to alfisol but has sparse cations. May contain
laterite or bauxite horizons. Forest and woodland soil.
Oxisol — Thick well-differentiated clayey soil, highly oxidized
and red lower horizons, and no unweathered material. May
contain laterite or bauxite horizons. Rainforest soil.

alluvial architecture (Allen, 1974; Wright and Marriott, 1993;


Willis and Behrensmeyer, 1994; McCarthy and Plint, 1998;
Retallack, 2001). However, some caution is necessary, because
soil maturity (pedofacies) is controlled not only by deposition red color
rate. In sequences of floodplain deposits where paleosoils vary in
their maturity, it is necessary to determine whether these varia- brown color
tions are local or regional, and whether they are related to
variations in deposition rate or other factors. This is no easy task.
FIG. 43.—Soil types of the US Soil Conservation Service (from
FOSSILS IN FLUVIAL DEPOSITS Retallack, 1997).

Preservation of Hard Parts


Trace Fossils
The hard parts of animals and plants are common in fluvial
deposits, depending upon the stage in evolution of organisms on Trace fossils (ichnofossils) are also very common in fluvial
Earth. Vertebrates include fish, amphibians, reptiles, and mam- deposits from most sub-environments (reviews in Pemberton et
mals. Invertebrates include molluscs (bivalves, gastropods), al., 1992; Retallack, 2001; Hasiotis, 2002; Bridge, 2003). They
arthropods (crustaceans and insects), and annelids (earthworms, record the dwelling, burrowing, and surface movement of a wide
nematodes). Plant remains (e.g., plant axes, roots, leaf impres- variety of organisms, including worms, arthropods (insects, crus-
sions, pollen and spores) are present in abundance in Devonian taceans), molluscs, and vertebrates. Plant root casts occur in
and younger sediments (Fig. 46; Gensel and Edwards, 2001). many fluvial deposits, with the exception of deposits formed
These body fossils can occur in situ or as transported remains. below the low-water level in channels and lakes. Until the 1980s,
Some fossils are transported only a short distance, so as to remain the only formal continental ichnofacies was the Scoyenia ichnofa-
within their local habitat. Transported plants and bones are cies (e.g., Ekdale et al., 1984; Frey et al., 1984), comprising dwell-
normally sorted by size, shape, and density, oriented by the ing, feeding, and crawling traces produced mainly by arthropods
current, and abraded because of contact with other hard objects in ephemeral lakes and floodplains. The full diversity and impor-
during transport. The abundance and diversity of fossils varies tance of continental trace fossils is now being realized and docu-
greatly within different fluvial sub-environments (Behrensmeyer mented. A new Mermia ichnofacies has been proposed for lacus-
and Hook, 1992). Abundance and diversity are related to that of trine environments (Buatois and Mángano, 1995). A Coprinisphaera
the original ecosystem, to the nature of transport and burial, and ichnofacies (Genise et al., 2000), named after dung beetle nests, is
to postdepositional modification of the material. Good preserva- an ichnofacies of insect traces (bees, wasps, ants, beetles, termites)
tion is generally favored by rapid deposition and burial in topo- in paleosols. Genise et al. (2000) suggest that this ichnofacies
graphically low sites, and by negligible chemical or biological should subsume the previously defined Termitichnus ichnofacies
degradation in the burial environment. (Smith et al., 1993; Hasiotis and Dubiel, 1995), an assemblage
130 JOHN S. BRIDGE

Channel Levee or Crevasse Splay Floodbasin Lake

low water table

Deposition rate, grain size, and drainage decreases

Soil maturity increases

Oxidized, leached horizon gleyed horizon thickening relative


over gleyed horizon to oxidized, leached horizon

CaCO3 glaebules above Fe glaebules increasing


low-stage water table relative to CaCO3 glaebules

Compound soils cumulative soils

FIG. 44.—Idealized catena associated with floodplain environments.

dominated by termite nests. In contrast to definition of these new table varies seasonally, the traces made by these different organ-
ichnofacies, Hasiotis (2004) has constructed ichnocoenoses for isms may become superimposed. The content of soil moisture
alluvial, lacustrine, and transitional marine environments (e.g., and the elevation of the water table are controlled by climate and
Fig. 47). Ichnocoenoses are assemblages of trace fossils that reflect by position on the floodplain relative to water bodies such as
biological communities (above and below ground). These bio- rivers and lakes. In arid climates, the amount of soil moisture is
logical communities (hence ichnocoenoses) vary with low, as is the average height of the water table. In moving from the
subenvironment and climate. edge of a lake or river onto a dry floodplain, the amount of soil
Continental trace fossils tend to be vertically zoned (tiered), moisture decreases, as does the mean elevation of the water table.
reflecting the soil moisture and level of the water table. For This results in an increase in biodiversity, biotic exchange, bur-
example, most insects and earthworms live above the water table, rowing depth, and degree of tiering from the water body to the
crabs and crayfish occupy a zone near the water table, and some floodplain (Fig. 47). For example, insect nests (termites, soil bees,
organisms (oligochaete worms, molluscs, water-loving insects, ants) tend to be deeper and larger if there are seasonal extremes
shrimp) live below the water table. As the elevation of the water of rainfall. Insect nests are smaller and shallower where the

well drained soils with


eluvial–illuvial horizons
he oxidized
m
at
ite
hematite –
illuvial horizons

FIG. 45.—Drying-out sequences in paleosols (from Wright, 1999).


FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 131

A C

FIG. 46.—Paleozoic fluvial fossils. A) sandstone cast of tree-trunk (0.1 m scale). B) Oriented plant axes (centimeter scale). C, D) Bivalve
escape burrows. E) Common type of meniscate burrow, possibly formed by arthropods. (A) to (E) are from Devonian of
northeastern USA or Canada. Photos from Steve Hasiotis.
132 JOHN S. BRIDGE

F G

H I

FIG. 46 (continued).—Mesozoic to Tertiary fluvial fossils. F) Mammal burrows. G) Termite nest, about 1 m across. H) Crayfish burrow.
I) Dung beetle nest with ball. Photos from Steve Hasiotis.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 133

Traces absent where water table is deep Wet-season water-table level Dry-season water-table level

FIG. 47.—Alluvial trace-fossil model, based on the Jurassic Morrison Formation (Hasiotis, 2004). Legend to symbols follows. AMB,
adhesive meniscate burrows; An, Anchorichnus; At, ant nests; Bv, Bivalve traces; Ca, Camborygma (crayfish burrows); Ce, Celliforma
(bee cells and nests); Cl, Cylindrichum; Cp, Coprinispheara (dung beetle traces); F, Fuersichnus (insect larvae burrows); G, Gastropod
trail; Hb, Horizontal burrow; Jb, J-shaped burrow; O, Ornithopod and theropod tracks; P, Planolites; Rh, Rhizoliths; Sa, Sauropod
tracks; So, Scoyenia; St, Steinichnus (beetle traces); Tm, Termite nests; T/Rh, termite nests in rhizoliths; Vb, quasi-vertical burrows;
Vtb, Vertebrate burrows; Wp, wasp nests/cocoons; Yt, Y-shaped burrows.

substrate is either dry or wet. Freshwater trace fossils reflect casts in channel deposits can give (along with desiccation cracks
seasonal changes in water depth and salinity, and are thus and falling-stage drainage features) an indication of the low
controlled by climate also. Trace-fossil abundance and soil matu- water level, and hence the range of flow stage in the paleoriver.
rity increase as deposition rate decreases, because of the time For example, root casts and desiccation cracks in the deepest part
available for bioturbation and pedogenesis. of a paleochannel are a clear indication that it was seasonally
ephemeral.
Fossils in Different Fluvial Environments
Abandoned Channels.—
River Channels.—
Abandoned channels support a prolific ecosystem and expe-
Much of the information included here on fossils in different rience relatively high deposition rate, leading to abundantly
fluvial environments comes from the reviews of Behrensmeyer preserved fossils. Plant assemblages tend to be derived mainly
and Hook (1992) and Hasiotis (2002). Plant material is commonly from the channel margins, and peat can be preserved in this
preserved in river channel deposits (e.g., logs, twigs, fruits, seeds, waterlogged environment. Common invertebrates are molluscs
leaves, pollen and spores). Invertebrate fossils are mainly unionid (bivalves, gastropods) and arthropods. Tetrapod and other aquatic
bivalves and freshwater oysters. Vertebrates include tetrapods vertebrate bones may be common. As the channel is filled, the
and other aquatic animals in various stages of disarticulation. fossils may indicate a trend from aquatic to terrestrial fauna. The
Most of the plant and bone material is transported, if not very far. fauna is usually not transported, and may be associated with
Trace fossils in river channel deposits include the dwelling, mass death due to aridity or predation. Trace fossils include the
locomotion, and escape burrows of bivalves, arthropod feeding dwelling burrows of bivalves, arthropod feeding and resting
burrows, arthropod walking traces, and plant root casts. Trace burrows, and walking traces. Root traces increase in abundance
fossils tend to be concentrated in upper-bar deposits. Plant root upwards in channel fills.
134 JOHN S. BRIDGE

Levees and Crevasse Splays.— are rare, but aquatic and tetrapod vertebrates and flying ani-
mals are common.
Levee and crevasse-splay deposits contain abundant plant
roots and transported plant material (wood fragments, leaves, Changes in Fluvial Fossils Over Time
pollen and spores). Chemical and biological degradation of plants
follows burial in these well drained, oxidizing environments, but Change in the preservation of fluvial fossils over time has
casts of these plants commonly survive. Disarticulated, trans- been influenced not only by evolution of life on land but also by
ported tetrapod and aquatic vertebrates also occur, as do mol- change in climate and tectonic activity that has changed the land
luscs. Bivalve escape burrows and arthropod dwelling burrows area and relief. Change in climate and tectonic activity control
(e.g., crayfish, beetles, termites, wasps) are common in levees and change in depositional environments, and hence in the types of
crevasse splays, and vertebrate burrows also occur. organisms present and their modes of preservation. It is com-
monly difficult to discern whether an evolutionary sequence of
Flood Basins.— organisms is related to changing organisms or changing deposi-
tional environments. There are many more fluvial fossils avail-
Well-drained floodplains (including abandoned channel able in Quaternary and Tertiary sediments than in Mesozoic and
belts) are sites of intense bioturbation and pedogenesis. Organic Paleozoic sediments. This is partly due to the evolution of life on
debris is readily oxidized, and acidic soils lead to dissolution of land, and partly due to the greater exposed volume of the young-
shell and bone. Some organic remains (e.g., logs, in situ tree est deposits. Some important stages in the evolution of life on
stumps, and roots) can be preserved as casts in pedogenic land relating to fossil preservation are given in Behrensmeyer
calcium carbonate or as casts of sandstone or mudstone. Micro- and Hook (1992), Buatois et al. (1998), Driese et al. (2000), Driese
floras are rare because of chemical degradation. Articulated and and Mora (2001), and Shear and Selden (2001).
disarticulated bones of tetrapods and fish are uncommon, and
such assemblages show evidence of surface weathering, carni- RIVER DIVERSIONS (AVULSIONS)
vore damage, or trampling, and are enriched in the most resis- ACROSS FLOODPLAINS
tant animal parts. Invertebrates such as molluscs (gastropods
and bivalves) and crustaceans are uncommon, probably be- Observations of the Nature of Avulsion
cause of dissolution of their hard parts. Floodplain soils com-
monly contain root traces (especially rhizoliths), burrows, pel- Avulsion is the relatively abrupt shift of a channel belt from
lets and coprolites, and trackways. Very common (post-Paleo- one location to another on the floodplain (references in Bridge
zoic) trace fossils are vertical, branching, and chambered bur- and Leeder, 1979; Mackey and Bridge, 1995; Jones and Schumm,
rows of insects (bees, wasps, dung beetles, ants, termites; Figure 1999; Smith and Rogers, 1999; Berendsen and Stouthamer, 2001;
47). Insects and spiders produce vertical dwelling burrows, Stouthamer, 2001; Stouthamer and Berendsen, 2000, 2001; Bridge,
horizontal feeding burrows, and surface trails. Insects do not eat 2003; Slingerland and Smith, 2004). Evidence for avulsions comes
sediment or line burrows. Worms also produce burrow tubes in from studies of modern rivers and from abandoned channel belts
various orientations, but worms eat sediment, producing pel- on Holocene floodplains. Avulsions normally occur during floods,
lets, and line burrows with mucus. Trackways and burrows of although the high water levels required for avulsions may be
vertebrates (e.g., various kinds of rodents; Figure 47) occur, and created by downstream blockage of the channel associated, for
some burrows contain bones and coprolites. example, with ice, vegetation, or sediment. An avulsion may be
Wet, poorly drained floodplains (marshes, swamps, forests) initiated by enlargement of a channel on a crevasse splay or by
typically have chemically reducing conditions at and below the intersection of the main channel with a preexisting channel. The
surface, and are favorable sites for preservation of plant material new channel belt follows the maximum floodplain slope on its
(logs, in situ tree stumps, leaves, peat, seeds, microflora), roots, way towards the locally lowest part of the flood basin. Aban-
and insect parts. In coal-bearing sequences, tree-stump casts, doned channel belts may block the path of avulsing channels;
plant compressions, and roots are especially common. Articu- however, an abandoned or active channel belt may be taken over
lated and disarticulated vertebrate skeletons may be present (e.g., by the avulsing channel. The transfer of water discharge from the
fish, reptiles, amphibians). Vertebrate trackways are common old channel belt to the new channel, and the associated change in
near the shorelines of rivers and lakes. Molluscs and arthropods channel pattern, may occur over years to centuries. Inter-avul-
occur, and their burrows and trackways are common. sion periods for a given channel belt (defined as the period of
The fossils of floodplain lakes are similar to those of aban- activity of the channel belt minus the avulsion duration) range
doned channels. In oxygenated lakes, macroplant debris and from decades to thousands of years. Successive avulsions may be
stromatolites occur but are not very common. Pollen and spores, initiated from a specific section of a valley (i.e., nodal avulsion;
benthic phytoplankton, and charophytes are more common. Fig. 48), particularly in the case of alluvial fans and deltas (see
Oligochaete worms, molluscs (bivalves, gastropods), amphi- below). In other cases, successive points of avulsion may shift
pod crustaceans, ostracods, and insect larvae are very common. progressively upstream with decreasing avulsion period, until
Disarticulated fish and tetrapods are also common. Roots, bur- there is an abrupt down-valley shift in the location of the avulsion
rows, and trackways are very common in oxygenated lakes. points. In yet other cases, avulsing channel belts appear to move
Worm burrows are lined, vertical tubes that may branch. Bur- progressively in one direction across floodplains or fans (e.g., the
rows of insect larvae are U-shaped. Bivalve dwelling burrows Kosi River; Fig. 48).
are relatively large vertical chambers. Depth zonation of benthic
organisms occurs in a way similar to the sea (Hasiotis, 2002). Avulsion and Anastomosis.—
Suspension feeders tend to be more common in shallow water,
whereas deposit feeders are more common in deeper water. If the different channels in avulsive river systems coexist for
Low-oxygen lakes may have anoxic bottom conditions, leading a finite length of time, the river system can be classified as
to common preservation of logs, leaves, seeds, pollen, and distributive (divergent channel belts) or anastomosing (if chan-
spores. Benthic invertebrates and insects (hence ichnofossils) nel belts split and rejoin). Such river patterns are typical of
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 135

Effects of Base-Level Change on Avulsion.—


High deposition rate Low deposition rate
4 Inter-avulsion periods decrease (avulsion frequency increases)
6
2 as relative base level rises, because of the down-valley decrease in
5
1 river slope and sediment transport rate, which causes aggrada-
3 tion and growth of alluvial ridges (e.g., Törnqvist, 1994; Ethridge
3
et al., 1999). Stouthamer and Berendsen (2001) found an increase
in avulsion frequency with increased rate of base-level rise for the
8
Rhine–Meuse delta, but not an associated decrease in inter-
2
5 avulsion period. This means that the number of coeval channels
4 must increase during increasing rate of base-level rise. Another
1 9 effect of rising base level is that points of avulsion shift up-valley
7 as the sea transgresses over the land. Avulsion is apparently more
Low deposition rate High deposition rate
likely to be associated with crevasse channels and splays when
aggradation rate is high, and with reoccupation of preexisting
channels if aggradation rate is relatively low (Aslan and Blum,
High deposition rate High deposition rate 1999; Morozova and Smith, 1999; Stouthamer, 2001).
2 1
4 1
The response of rivers to base-level fall depends on factors like
3 2 the slope and sediment type of the exposed land areas, and the
6 rate of fall (Schumm, 1993; Wescott, 1993). If the slope of the
exposed surface is greater than the equilibrium channel slope,
there may be channel incision and/or the channel may become
more sinuous. Avulsion frequency may be expected to decrease
in this region, but not necessarily farther up valley. An upstream
5
avulsion may limit the amount of incision farther down valley
(Leeder and Stewart, 1996). If the slope of the exposed surface is
less than the equilibrium channel slope, there may be deposition
and/or the channel may become straight. Under these condi-
tions, avulsion frequency may be increased.

Effect of Climate Change on Avulsion.—


Low deposition rate
Changes in climate in the drainage basin can affect avulsion
FIG. 48.—Typical patterns of channel-belt avulsion, dependent by changing the discharge regimes and sediment supply to the
upon spatial variation in deposition rate and preexisting rivers. For example, during and immediately following glacial
floodplain topography. Channel belts are represented by periods, the magnitude and variability of water and sediment
lines. Numbers are locations of avulsion in chronological supply may be increased, resulting in higher avulsion frequen-
order. Full explanation in Bridge (2003). cies. This trend may be reversed during warmer periods. In
addition, changes in the base levels of lakes and the sea during
glaciations should affect avulsion frequencies (Morozova and
alluvial fans, crevasse splays, and deltas. These are environ- Smith, 1999), and ice dams could cause river diversions.
ments with high deposition rates and active growth of alluvial
ridges, producing the gradient advantages conducive to avul- Effects of Tectonism on Avulsion.—
sion (Smith and Smith, 1980; Smith, 1983; Smith et al., 1989;
Smith et al., 1998; Törnqvist, 1993, 1994: Makaske, 2001). The Tectonically induced changes in river and floodplain gradi-
water and sediment discharge (hence channel geometry) of ents may result in avulsions (reviews by Mackey and Bridge,
individual channel-belt segments may change in time. The 1995; Peakall et al., 2000; Schumm et al., 2000; Bridge, 2003).
channel segments in anastomosing river systems can have any Avulsion may occur as a direct response to an individual tectonic
type of channel pattern (e.g., meandering, braided; Fig. 14). event or (more likely) in response to a gradual, tectonically
Therefore, it makes no sense to have depositional models for induced change in floodplain topography. Avulsing channels
anastomosing river systems that are distinguished from models concentrate in areas of tectonic subsidence and avoid areas of
for braided or meandering rivers (as occurs in much of the tectonic uplift. However, if channels occupying subsided areas
literature on fluvial deposits). have a high aggradation rate, subsequent river diversions may be
away from the zone of maximum subsidence. Tectonically in-
Effects of Sedimentation Rate on Avulsion.— duced changes in river and floodplain gradients may also result
in aggradation and an increased probability of avulsion, or inci-
Avulsion frequency increases (inter-avulsion period decreases) sion and a reduced probability of avulsion.
with increasing channel-belt deposition rate (Mackey and Bridge,
1995; Heller and Paola, 1996; Ashworth et al., 2004). This is Theoretical Models of Avulsion
because high channel-belt deposition rate causes rapid growth of
the alluvial ridge above the floodplain, and attendant rapid Mackey and Bridge (1995) suggested that the probability of
increase in cross-valley slope of the alluvial ridge relative to its avulsion at a given location along a channel belt increases with
down-valley slope. Channel-belt aggradation rate can be influ- a discharge ratio (maximum flood discharge for a given year/
enced by base-level change, climate change in the hinterland, or threshold flood discharge necessary for an avulsion) and a slope
tectonic activity, as explained below. ratio (cross-valley slope at the edge of the channel belt/local
136 JOHN S. BRIDGE

down-valley slope of the channel belt). They assumed that high, produces nodal avulsions, which are characteristic of
avulsions are initiated during extreme discharge events when alluvial fans.
erosive power of the stream is greatest, and that a sufficient The Mackey–Bridge model predicts that avulsion periods
stream-power advantage exists for a new course to be estab- vary greatly depending on the stage of growth of alluvial ridges.
lished on the floodplain. The stream power and sediment trans- Sections of alluvial ridges that are well developed are associated
port rate per unit channel length are proportional to the dis- with short avulsion periods (say, decades or centuries), whereas
charge–slope product. Therefore, if the discharge–slope prod- newly formed channel belts may not experience an avulsion for
uct and sediment transport rate of an overbank flow can exceed on the order of a thousand years. Also, the obstruction to flow
that of the channel flow, sediment transport rate increases from caused by preexisting alluvial ridges may cause subsequent
the channel to the floodplain, and erosion and enlargement of a channel belts to be clustered preferentially on one side of the
crevasse channel is possible. The discharge in a growing cre- floodplain with a distinctive en echelon pattern (Fig. 48). This
vasse channel initially is less than that in the main channel. means that other parts of the floodplain distant from the active
Therefore, the overbank slope must be much greater than the channels experience relatively low overbank deposition rates for
channel slope in order for the sediment transport rate to increase extended periods of time, allowing soils to develop.
from the main channel to the floodplain. In the limiting case Tectonic tilting and faulting within the floodplain increase
where the discharges of the main channel and the developing avulsion probability locally, according to Mackey and Bridge
crevasse channel are equal, the slope in the crevasse channel (1995) (Fig. 49). Channel belts shift away from zones of uplift and
must exceed that of the main channel for the avulsion to pro- towards zones of maximum subsidence. However, if channel-
ceed. The water level in a flood basin must be lower than that in belt aggradation keeps pace with fault displacement or tilting,
the main channel to allow water to flow away from the main alluvial-ridge topography causes channels to shift away from
channel through a crevasse channel. When the water-surface areas of maximum subsidence. Although these predictions agree
elevations of the main channel and the flood basin are the same, broadly with data from modern rivers (e.g., Mike, 1975; Schumm,
there can be no such crevasse-channel flow. Therefore, crevasse- 1986; Alexander and Leeder, 1987; Leeder, 1993; Peakall, 1998,
channel enlargement can operate only during certain overbank Peakall et al., 2000; Schumm et al., 2000), data on the relationship
flood stages. Accordingly, it may take a number of overbank between tectonism and avulsion are insufficient to test model
flood periods for a crevasse channel to enlarge to the point of predictions in detail.
avulsion. Slingerland and Smith (1998) made the only analytical ap-
Mackey and Bridge (1995) used their avulsion probability proach to the cause of avulsion. Their model is based on simpli-
model to simulate avulsions where floodplain slopes vary in fied equations of motion for fluid and sediment applied to simple
space and time because of variations in deposition rate, and channel geometry. The crux of the model is that the suspended-
tectonic tilting and faulting within the floodplain. Down-valley sediment concentration at the entrance to a crevasse channel
increase in channel-belt deposition rate produces a down-valley leading from a deeper main channel is different from the equilib-
decrease in channel-belt slope but an increase in cross-valley rium concentration that should exist in the crevasse channel.
slope as the alluvial ridge grows, as is likely to happen during Then, depending on local hydraulic conditions, the crevasse
base-level rise. Under these circumstances, avulsion probability channel would deepen or fill with sediment until the equilibrium
increases through time, and is greatest in the down-valley part of sediment concentration is reached. A condition where the chan-
the floodplain, where channel-belt slopes are smallest but cross-
valley slopes are largest. The model predicts that avulsion is
initiated in the down-valley part of the floodplain and successive
avulsion locations shift up valley with a progressive decrease in 3
inter-avulsion period. This is due to gradual increase in avulsion 4
probability up valley of avulsion locations where growth of
alluvial ridges continues uninterrupted. New channel-belt seg-
ments down valley from avulsion locations have not had time to
5
aggrade significantly and develop alluvial ridges, and therefore 1
have low avulsion probabilities. After a finite number of avul-
sions stepping up-valley, the progressive decrease in channel-
belt slopes in the down-valley part of the floodplain causes
abrupt shift in the locus of avulsion to this location. Although
model results agree broadly with what is observed in nature 3 1
(Mackey and Bridge, 1995), the model does not take into account 4
2
the effects of the increased slope of the new channel as it leaves the
2
old channel. This steep slope would result in channel incision and
upstream retreat of a knickpoint in the vicinity of the point of
avulsion. Therefore, the probability of avulsion would be greatly
reduced immediately upstream of a recent avulsion location.
Down-valley decrease in channel-belt aggradation rate produces
a down-valley decrease in down-valley slope, but an increase
over time in down-valley slope at any point on the floodplain,
as is typical of alluvial fans and where base level is falling.
Avulsion probability decreases with time because overall down- FIG. 49.—Typical effect of faults and tilting on channel-belt avul-
valley channel-belt slope increases. However, avulsion prob- sion. Channel belts are indicated by lines. Numbers are loca-
ability is high in the up-valley parts of the floodplain, where tions of avulsion in chronological order. Downthrown sides
cross-valley slopes are increased by high aggradation rate. The of faults are indicated by triangles. Full explanation in Bridge
concentration of avulsions up valley, where deposition rate is (2003).
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 137

nel progressively deepens is taken as a criterion for avulsion. such movement is rare or equivocal in modern rivers. Peakall et
Avulsions are predicted to occur wherever the slope ratio exceeds al. (2000) suggested that gradual shift of channel belts occurs
about 5. Notwithstanding the importance of this first analytical instead of avulsion if tilt rates are relatively low. However, the
approach, the geometry and physical processes are treated at a only physical model for the effect of floodplain tilting on channel
very simple level. Bedload sediment transport is not treated, so that migration (Sun et al., 2001) predicts that a river could migrate
the model cannot explain avulsions in the many rivers that trans- towards or away from the down-tilted area, depending upon
port mainly bedload during floods. Slingerland and Smith (2004) flow characteristics. Avulsions are the most important means of
mention that theories for the stability of channel bifurcations moving channel belts around most natural floodplains.
(including treatment of bedload) in braided rivers (e.g., Bolla
Pittaluga et al., 2003) might be applied to prediction of avulsion. ALONG-VALLEY VARIATION IN
RIVERS AND FLOODPLAINS
Effects of Avulsions on Erosion and Deposition
River and floodplain geometry, water flow, sediment trans-
Diversion of a channel belt to a new area of flood basin may be port, erosion, and deposition vary along valley as tributaries join,
preceded by extensive development of crevasse splays (Smith et as water is lost by evaporation and/or infiltration, as geological
al., 1989). It has even been suggested that most floodplain depos- features such as faults and lava flows change valley slope, as
its may be formed during the period of crevasse-splay deposition valley width changes, and as rivers enter bodies of water such as
preceding an avulsion (Smith et al., 1989; Kraus, 1996; Kraus and lakes and seas.
Gwinn, 1997; Kraus and Wells, 1999). This is difficult to establish,
however, and crevasse splays can be deposited without having Along-Valley Variation of River Slope and Geometry
an avulsion associated with them.
Avulsions may also be recognized in ancient floodplain Long profiles of river channels and valleys (along-stream
deposits without actually observing the diverted channel de- variation in bed elevation) are commonly considered to be con-
posits (e.g., Elliott, 1974; Bridge, 1984; Behrensmeyer, 1987; cave upwards and fitted by an exponential curve (recent reviews
Farrell, 1987, 2001). For example, initiation of an avulsion may by Sinha and Parker, 1996; Morris and Williams, 1997, 1999a,
be recorded in floodplain deposits by an erosion surface over- 1999b; Rice and Church, 2001). However, long profiles are not
lain by relatively coarse-grained deposits (associated with a always concave upwards, especially in tectonically active areas
major overbank flood) and overlying deposits may be different (Leeder, 1999). The commonly observed down-valley decrease in
from those that were deposited prior to the initiation of the mean grain size of channel sediment is due mainly to down-
avulsion. If the channel belt moved to a more distant location on valley decrease in bed slope and bed shear stress within the river
the floodplain than hitherto, the new flood-generated stratasets system, such that the coarsest grains are progressively lost in the
on the floodplain may be thinner and finer grained and be down-valley direction (review in Bridge, 2003, p. 70). Down-
associated with a different flow direction. If deposition rate is valley reduction in bed sediment size due to progressive abrasion
decreased, soils may become more mature (Leeder, 1975; Bown is of minor importance.
and Kraus, 1987; Kraus, 1987). If the channel belt moved closer Downstream increase in channel-forming discharge of wa-
to a given floodplain area, the new overbank flood deposits ter and sediment, due to joining tributaries, causes a down-
could be thicker and coarser grained, and coarsening-upward stream increase in bankfull channel width and depth. For a
sequences may occur because of development and prograda- given sediment supply, width increases more than depth as
tion of levees and crevasse splays (e.g., Elliott, 1974; Farrell, discharge increases, such that width/depth ratio might be ex-
1987, 2001; Perez-Arlucea and Smith, 1989). However, meter- pected to increase down valley with discharge. Related to this,
scale overbank sequences that fine upwards or coarsen up- channel pattern might be expected to change downstream from
wards can also be produced by progradation or abandonment single channel to braided. Actually, it is quite common to see the
of different levees and crevasse splays from a fixed channel belt, opposite, because, although discharge increases downstream,
or by regional changes in sediment supply and deposition rate channel slope and mean grain size may decrease (Bridge, 2003,
(see previous section). p. 297).
Avulsion by channel reoccupation is very difficult to recog-
nize in the stratigraphic record. It has been claimed (Mohrig et al., Effects of Tectonism on Along-Valley Variation
2000; Stouthamer, 2001) that channel reoccupation results in of Rivers and Floodplains
relatively thick, multistory channel belts and multiple levees, but
these stratigraphic features can easily be produced by episodic Tectonic activity affects the slopes of rivers and floodplains,
deposition in a single channel belt. The only conclusive evidence and their supply of water and sediment, over a range of spatial
for channel reoccupation is preservation of deposits indicating and temporal scales. For example, periodic activity of a fault
long time periods (such as mature soils) between the deposits of crossing a river valley may result in a local change in valley slope,
superimposed channel belts. river diversion, and subsequent change in channel pattern over a
period of several hundreds or thousands of years. On a much
Non-Avulsive Shift of Channel Belts Across Floodplains larger scale, the whole river system may be affected by tectonic
activity for millions of years if it is located in a tectonically active
It has been suggested that channel belts are capable of gradual mountain belt with adjacent sedimentary basin (e.g., Basin and
migration across their floodplains by preferential bank erosion Range Province, Himalayas and Indo-Gangetic alluvial basins).
along one side of the channel belt and net deposition on the other Relatively short-term, local tectonic influences are discussed
side (Allen, 1965, 1974; Coleman, 1969; Thorne et al., 1993; Peakall here, and longer-term, regional tectonic influences are discussed
et al., 2000). Such migration is possibly a response to across-valley below.
tilting of the floodplain, and gives rise to so-called asymmetrical Local tectonic activity includes the effects of movement on
meander belts (Alexander and Leeder, 1987, 1990; Leeder and relatively small folds and faults that directly influence land
Alexander, 1987; Alexander et al., 1994). However, evidence for topography. Tectonic activity can also act indirectly by control-
138 JOHN S. BRIDGE

ling base level and the slope of rivers entering standing water. diversion is most likely in areas of reduced down-valley slope.
Tectonic and volcanic activity can cause blockages of river Relatively large cross-valley slopes therefore encourage diver-
valleys (creating lakes and local changes in slope) with lava, sion of the river during flood periods in the direction of maximum
volcanic ash, and debris flows. Changing slopes within alluvial valley slope (Fig. 49; Mackey and Bridge, 1995). In this case, the
valleys can result in diversions of rivers and changing channel river may become anastomosing if it cannot be diverted around
patterns. Changing topography of the floodplain relative to the the topographic obstruction. When a river crosses an active strike
groundwater table can influence floodplain flow and sediment slip fault, it may be offset by the fault. Alluvial fans on the
transport, and the development of soils. These topics were downthrown sides of such faults may be separated laterally from
reviewed recently by Schumm et al. (2000) and Burbank and the stream that provided the fan sediment (Burbank and Ander-
Anderson (2001). son, 2001).
Activity of a fault or fold may result in a depression or ridge Along-valley tectonic structures such as faults result in tilting
with its axis approximately normal to the valley (Fig. 50). Along- of the floodplain in the across-valley direction. A common re-
valley slope decreases upstream of an uplifting ridge axis and sponse of rivers to cross-valley tilting is periodic diversion to-
increases downstream of the axis. The opposite occurs in the case wards the down-tilted area, producing asymmetrical channel
of a subsiding depression. The short-term response to a slope belts where the active channel occupies the lowest part of the
decrease might be a decrease in river sinuosity and increase in floodplain, and abandoned channel belts occur on the up-tilt side
degree of braiding, but an increase in river sinuosity and decrease (Figs. 49 and 51; Coleman, 1969; Mike, 1975; Alexander and
in degree of braiding where slope is increased (e.g., Ouchi, 1985). Leeder, 1987; Leeder 1993; Mackey and Bridge, 1995; Peakall,
In regions of slope decrease, the floodplain may become perma- 1998; Holbrook and Schumm, 1999; Peakall et al., 2000; Schumm
nently swampy, or the frequency and height of overbank flood- et al., 2000). The low area of the floodplain is also flooded
ing may increase. In other cases, a former floodplain lake may be preferentially and suffers increased deposition rate associated
reduced in level. Deposition is expected where slope is decreased, with both channel belts and increased overbank floods. If depo-
and erosion (and possibly terrace formation) is expected where sition rate temporarily exceeds tectonic subsidence rate, rivers do
slope is increased. If there is an abrupt transition from the zone of not always occupy the zone of maximum subsidence. Increased
erosion and river incision to a zone of deposition downstream, deposition rate in down-tilted sides of floodplains tends to re-
the deposit may be in form of an alluvial fan (see below). duce the development stage of soils, whereas the up-tilted sides
The response of the river to such active tectonic deformation experience lower deposition rate and increased soil maturity.
of the land surface depends on the rate of deformation relative to
the rate of river erosion and deposition, in turn related to stream Alluvial Fans
power and bank erodibility. If the rate of erosion and deposition
is sufficiently large, along-stream changes in slope arising from Definition, Occurrence, and Geometry.—
tectonic activity are reduced, and the river may tend to its former
state. However, if rate of erosion and deposition is not sufficient Alluvial fans and deltas (reviewed recently by Miall, 1996;
to remove tectonic topography, river diversion may occur. River Leeder, 1999) have distinctive plan shapes and distributive to

uplift subsidence

deposition erosion deposition erosion deposition erosion

T
T

braided river

T
T
meandering river

FIG. 50.—Effect of tectonism on along-stream change in channel patterns (based partly on Ouchi, 1985). Erosion occurs where slope
is increased, and where sediment supply is reduced because of upstream deposition, producing terraces (T), decrease in degree
of braiding, or increase in sinuosity of meandering rivers. Deposition occurs where slope is decreased, and downstream of erosion
zones, producing increase in braiding and local avulsion.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 139

FIG. 51.—Asymmetrical channel belt, Senguerr River, Argentina. Channel belt occupies one side of floodplain, as a result of tectonic
tilting. Floodplain is bordered by dissected uplands (background) or a terrace margin (foreground).

anatomosing channels bordered by floodplains. Deltas build into hundred kilometers, increasing with the supply rate of water and
standing bodies of water. If alluvial fans build into standing sediment from the catchment. Fan area increases with catchment
bodies of water, they are referred to as fan deltas. Deltas range in area approximately linearly, although the exact relationship de-
size from small lobes where cross-bar channels enter slow-mov- pends on climate, catchment geology, and deposition rate rela-
ing water in bar-tail areas, to lacustrine deltas on floodplains, to tive to the basin subsidence rate (Leeder, 1999). Fans tend to be
the coastal deltas of major rivers. Alluvial fans can be small more or less evenly spaced and laterally coalesced along faulted
crevasse splays on floodplains adjacent to channels, or the well- mountain fronts, and their areas may increase along the fault as
known fans that occur at the margins of small, fault-bounded the amount of throw increases (Gawthorpe and Leeder, 2000).
valleys, or megafans (e.g., the Kosi fan), which occur where major Although the channel system on fans appears distributive to
rivers flow from mountain ranges onto broad alluvial plains. The anastomosing, not all channels are active at once. Diversion of
term terminal fan has been used for fans in arid areas where flood flows among different channels (avulsion) is common. The
flowing water percolates into the ground before reaching beyond trunk stream at the apex of the fan is commonly entrenched, as
the fan margins. Alluvial fans are commonly classified according discussed below. In humid climates, it is normal for at least one
to their surface slopes and the relative importance of sediment of the channels to be perennial and to continue beyond the fan,
gravity flows, sheet floods, and stream flows in forming them. but this is not the case in arid climates where the channels are
Alluvial fans occur in every climate where a river course ephemeral and the fans are “terminal”. In humid fans (e.g., the
passes sufficiently rapidly from an area of high slope to one of low Kosi fan), some of the smaller channels may originate on the fan
slope. The abrupt change of slope results in an abrupt down- surface from groundwater springs instead of originating in the
stream decrease in bed shear stress and sediment transport rate, hinterland.
which leads to localized deposition. Alluvial fans commonly The long profile of fans is normally concave upwards, i.e.,
occur adjacent to fault scarps, and the preservation of fan deposits slope decreases down fan. Average slope generally decreases as
is enhanced by the subsidence of the hanging wall. Alluvial fans supply of water and sediment increase, and as grain size of
are well known in poorly vegetated, arid areas where infrequent sediment supplied decreases. As slopes increase to on the order
violent rainstorms cause high sediment loads. However, alluvial of 10-2 to 10-1, sediment gravity flows (debris flows and mud-
fans also occur in humid and periglacial conditions. Alluvial fans flows) become more common. Channelized debris flows are
may pass downstream into alluvial plains, tidal flats, beaches, associated with distinct levees and terminate in lobate deposits.
perennial lakes, eolian dune fields, or playa lakes. The locally rapid Transverse profiles of fans are convex upward. Fan geometry is
deposition and fixed supply of water and sediment give rise to the complicated by incision of fan channels and formation of a new
fan shape (segment of a cone cut downward from its apex; Fig. fan downstream of the incised fan and by trimming of the fan toe
52). Fan radius varies from hundreds of meters to more than a by an axial river.
140 JOHN S. BRIDGE

FIG. 52.—Geometry and sedimentary processes of A) debris-flow-dominated and B) streamflow-dominated alluvial fans (from
Leeder, 1999).

Flow and Sedimentary Processes.— Channel diversions (avulsions) are common on alluvial fans
during floods, and occur following a period of aggradation near
River channels on fans are braided if there is a large supply of the fan apex (e.g., Schumm et al., 1987; Sun et al., 2002). This local
water and sediment from the hinterland. Groundwater-fed fan aggradation causes the transverse slope of the fan apex to in-
channels are generally smaller and may have a different channel crease relative to the downstream slope, thereby facilitating
pattern (e.g., meandering). In arid fans, channels decrease in size avulsion. Once avulsion is initiated, the new channel is incised
down-fan because water infiltrates or may become a sheet flood. into the fan surface, with the maximum incision occurring near
As the slope or depth of water decreases down-fan, grain size of the fan apex where the new course is steepest (fan-head entrench-
surface sediment decreases. Overbank areas have channelized ment). A depositional lobe forms downstream of the entrenched
flow and sheet floods. Debris flows, grain flows, and mudflows channel. As fan aggradation proceeds, the entrenched course of
are particularly common in channelized apex areas, where slopes the new channel may start to aggrade by backfilling, starting from
are steepest. the lower fan region (Fig. 53). Eventually, the fan-head trench
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 141

A Basin outlet B because it is not possible to observe directly the processes of


development of alluvial architecture, it is necessary to use models
Trench
to interpret and predict alluvial architecture. Most quantitative
models of alluvial architecture (reviewed by Bryant and Flint,
1993; Koltermann and Gorelick, 1996; North, 1996; Anderson,
1997; Bridge, 2003, in press) are either process-based (process-
Deposition imitating) or stochastic (structure-imitating), as seen below.

Controls on Alluvial Architecture

C D Alluvial architecture is controlled primarily by: (1) the geom-


etry and sediment type of channel belts and floodplains; (2) the
rate of deposition or erosion in channel belts and on floodplains;
(3) local tectonic deformation within the alluvial valley; and (4)
the nature of channel-belt movements (avulsions) over flood-
plains. These intrinsic (intrabasinal) controlling factors are in turn
controlled by extrinsic (extrabasinal) factors such as tectonism,
climate, and eustatic sea-level changes. For example, the geom-
etry and sediment type of channel belts and floodplains are
controlled by water and sediment supply (rate and type), which
are in turn controlled by the source rocks, topographic relief,
E F climate, and vegetation of the drainage basin. Long-term deposi-
tion in alluvial valleys is due to long-term decrease in sediment
transport rate in the down-flow direction, which can be accom-
plished by increasing sediment supply up-valley (by tectonic
uplift, climate change, or river diversions) and/or by decreasing
sediment transport rate down valley (by flow expansion associ-
ated with tectonic subsidence or base-level rise). Subsequent
erosion depends upon increasing sediment transport in the down-
flow direction, such as caused by basin uplift or base-level fall.
FIG. 53.—Evolution of experimental alluvial fans (from Schumm Local tectonic deformation in alluvial valleys can cause local
et al., 1987). A) Fan-head channel entrenchment leads to changes in channel and floodplain geometry and location, depo-
deposition of a lobe downstream. B) Deposition zone moves sition, and erosion. Channel-belt movements across floodplains
upstream and channel is backfilled. C) Channel filled and are influenced by the severity of floods and the development of
deposition at fan head increases slope here. D) Channel cross-floodplain slopes associated with alluvial-ridge deposition
moves to new area of fan and is entrenched. E, F) as for Parts and local tectonic deformation. The intrinsic controls on alluvial
B and C. architecture will be examined first, followed by the extrinsic
controls.

becomes aggraded, and the conditions become favorable for Process-Based (Process-Imitating)
another channel avulsion. It is possible that fan-head aggrada- Models of Alluvial Architecture
tion, avulsion, and entrenchment are associated with pulses of
sediment supply from the hinterland, perhaps associated with The earliest hypothetical models of alluvial architecture were
episodic tectonism or climate change. qualitative and essentially two-dimensional (Allen, 1965, 1974).
Allen (1965) hypothesized that low-sinuosity, single-channel and
ALLUVIAL ARCHITECTURE AND ITS CONTROLS braided channels migrated rapidly (swept) across their flood-
plains, leading to alluvium dominated by sheet-like to lenticular
Alluvial Architecture channel deposits (Fig. 55). In contrast, single-channel, high-sinu-
osity streams were taken to migrate within well-defined meander
Extensive accumulations of fluvial deposits in sedimentary belts that experienced periodic avulsion, leading to ribbon-like
basins formed over millions of years normally show distinctive channel belts were set in a relatively high proportion of fine-
spatial variations in the mean grain size, geometry, and propor- grained alluvium (Fig. 55). These hypotheses are actually incor-
tion and spatial distribution of channel-belt and floodplain de- rect, but Allen’s (1965) models were useful for stimulating more
posits (referred to as alluvial architecture by Allen, 1978). These sophisticated quantitative approaches. Allen (1974) elaborated
spatial variations occur over a range of scales (e.g., vertical on his earlier models by considering the effects on alluvial archi-
sequences of strata may be tens to hundreds of meters thick; Fig. tecture of climate and base-level change, degradation, and differ-
54), and regional unconformities (erosion surfaces that extend ent modes of channel migration. More recent qualitative, 2-D
laterally for hundreds of kilometers) may underlie these se- models predict how alluvial architecture is dependent on changes
quences of strata. Definition of alluvial architecture requires in aggradation rate and valley width during change in relative sea
extensive exposures, and/or high-resolution (preferably three- level (Fig. 66; e.g., Shanley and McCabe, 1993, 1994; Wright and
dimensional) seismic data, and/or many closely spaced cores or Marriott, 1993).
borehole logs, and accurate age dating. Because such data are Leeder (1978) developed the first quantitative, process-imi-
commonly lacking or incomplete, it is necessary to “fill in” three- tating model of alluvial architecture. Channel-belt and flood-
dimensional space in order to produce a complete (and hypo- plain deposits were modeled within a single cross-valley sec-
thetical) representation of alluvial architecture. Furthermore, tion. Channel-belt width, maximum channel depth, and flood-
142 JOHN S. BRIDGE

FIG. 54.—Large-scale sequence of fluvial deposits from the Miocene Siwaliks of northern Pakistan. Sandstone-body (channel-deposit)
proportion decreases, increases, and decreases upwards over 2 kilometers of strata, but also varies over 100–200 meters and tens
of meters. The photos show zones of high and low sandstone-body proportion.

plain width were specified. The channel belt was allowed to tion rate with distance from channel belts (Fig. 56), and predict-
move by avulsion across the floodplain as aggradation (balanc- ing the width and thickness of channel sandstone–conglomer-
ing subsidence) continued. The period of time separating avul- ate bodies comprising single or connected channel belts (Fig.
sions and the location of avulsing channels in the floodplain 34). Channel-deposit proportion and sandstone-body width
section were chosen randomly within defined limits. Leeder and thickness increase as bankfull channel depth and channel-
defined 2-D measures of the proportion of channel-belt deposits belt width increase, and as floodplain width, aggradation rate,
and their degree of connectedness, which depend upon aggra- and inter-avulsion period decrease. Channel-deposit propor-
dation rate, avulsion frequency, channel-belt cross sectional tion and connectedness also increase in locally subsided areas of
area, and floodplain width. Allen’s (1978, 1979) later approach floodplain. These models show that the proportion of channel
followed Leeder’s closely, but Allen added a function to allow deposits (net-to-gross) has nothing to do with whether the river
diverted channel belts to avoid high floodplain areas underlain channel is meandering or braided, as suggested in Allen’s early
by preexisting channel belts. The 2-D approach was extended models. These quantitative 2-D models have been tested against
(Bridge and Leeder, 1979; Bridge and Mackey, 1993a, 1993b) by limited field data (e.g., Leeder et al., 1996; Mack and Leeder,
considering also the effects on alluvial architecture of compac- 1998; Peakall, 1998; Törnqvist and Bridge, 2002) and have been
tion, tectonic tilting of the floodplain, and variation of aggrada- used (and misused) widely to interpret and model alluvial
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 143

plain deposition and erosion associated with tectonism; (2) tribu-


taries and downstream increases in channel-belt size; (3) wide-
spread erosion (degradation) of channel–floodplain systems; and
(4) diversion of a channel belt into a preexisting channel. How-
ever, model development is under way, and has been greatly
aided by new data from Holocene fluvial and deltaic settings
(reviewed by Bridge 2003). The component models for aggrada-
tion, avulsion, and the development of channel belts following
avulsion have been greatly improved (details in Karssenberg et
al., 2003). New channel belts are formed by channel bifurcation,
and multiple channel belts may develop and coexist. In some
cases, a preexisting channel belt is abandoned in favor of another
one, and an avulsion occurs. Degradation of the channel and
floodplain are treated using a diffusion–advection approach,
such that upstream migration of knickpoints and formation of
incised channels and terraces can be simulated. The effect of
cyclic variation in degradation and aggradation on avulsion and
alluvial architecture can be simulated (allowing a link to se-
quence-stratigraphic models).
Models of sediment deposition that are based on solution of
the fundamental equations of motion of water and sediment are
referred to as sediment routing models. Engineers apply these
models to relatively simple flow and sedimentation problems in
FIG. 55.—Qualitative alluvial architecture models of Allen (1965). modern rivers and floodplains. Sedimentologists have also used
The braided-river model does not show an adjacent flood- them to explain phenomena such as downstream fining of bed
plain, and depiction of braided-river deposits is unrealistic.
The low-sinuosity-river model indicates erroneously that chan-
A
nels sweep gradually across their entire floodplains.

architecture (references in Bridge, 2003, p. 334). However, 2-D


models are unable to realistically simulate down-valley varia-
tions in the location and orientation of individual channel belts.
This is possible only with 3-D models.
Mackey and Bridge (1995) developed the first 3-D process-
imitating model of alluvial architecture. The floodplain contains
a single active channel belt (Fig. 57). Changes in floodplain
topography are produced by spatial and temporal variation of B
channel-belt and floodplain deposition rates, by compaction, and
by tectonism. The location and timing of avulsions are deter-
mined by local changes in floodplain slope relative to channel-
belt slope and by flood magnitude and frequency (discussed
above). The diverted channel follows the locus of maximum
floodplain slope. Major differences between this model and the 2-
D models are the treatment of avulsion location and period as
dependent variables, and constraints on the location of avulsing
channels by the points of avulsion and topographic highs on the
floodplain. The behavior of the avulsion model was discussed
above. Avulsions occur preferentially where there is a decrease in
channel-belt slope and/or an increase in cross-valley slope that C
may be related to spatial variations in deposition rate and/or
tectonism and/or base-level change. Evolution of alluvial ridges
over time in different parts of the floodplain greatly influences the
timing and location of avulsions. This may result in sedimentary
sequences that increase upwards in channel-deposit proportion
and connectedness, capped with overbank deposits with well-
developed soils. Such sequences may take on the order of 103 to 105
years to form, comparable to cycles attributed to tectonism or
climate change. Predictions of the Mackey–Bridge model agree
with the somewhat limited data from modern rivers, and the
model has been applied to interpreting and predicting the alluvial FIG. 56.—Examples of quantitative 2D alluvial architecture model
architecture of ancient deposits (references in Bridge, 2003, p. 336). of Bridge and Leeder (1979), showing effect on alluvial archi-
The Mackey–Bridge model has many shortcomings, and does tecture of A, B) varying channel-belt width and C) tectonic
not consider: (1) changes in channel pattern and channel-belt tilting. Channel belts are yellow blocks, and lines are flood-
width, formation of floodplain lakes, and channel-belt and flood- plain surfaces at time of avulsion.
144 JOHN S. BRIDGE

FIG. 57.—Quantitative, 3D alluvial-architecture models from Mackey and Bridge (1995) and Karssenberg et al. (2001). Colored objects
are channel belts.

material in rivers and large-scale sorting of placer minerals such “Conditioned simulations” begin by placing objects such that
as gold (e.g., Van Niekerk et al., 1992; Vogel et al., 1992; Bennett their thickness and position correspond with the available well
and Bridge, 1995b; Robinson and Slingerland, 1998a, 1998b; data. Then, objects are placed in the space between wells until the
Robinson et al., 2001). Koltermann and Gorelick (1992) used a required volumetric proportion is reached. Objects are placed
simple sediment routing model to investigate how the large-scale more or less randomly, although arbitrary overlap and repulsion
patterns of deposition in alluvial fans are related to climatically rules may be employed to produce “realistic” spatial distribu-
controlled changes in water and sediment supply, tectonism, and tions of objects (Fig. 58).
base-level change. However, sediment-routing models have not Continuous stochastic models have been used mainly to
yet been used to simulate alluvial architecture, but they probably simulate the spatial distribution of continuous data such as
hold out most hope for rational simulation of fluvial deposition permeability, porosity, or grain size. With these models, a param-
and erosion. eter value predicted to occur at any point in space depends on its
value at a neighboring site. The conditional probabilities of
Prediction of Alluvial Architecture of occurrence are commonly based on an empirical semivariogram.
Subsurface Deposits: Stochastic Models These approaches have been modified to predict the distribution
of discrete facies by using indicator semivariograms and simu-
The most common approach to predicting the architecture of lated annealing (Fig. 59; e.g., Johnson and Dreiss, 1989; Bierkens
subsurface fluvial reservoirs and aquifers (discussed in Bridge and Weerts, 1994; Deutsch and Cockerham, 1994; Seifert and
and Tye, 2000; Bridge, 2001) is to: (1) determine the geometry, Jensen, 1999, 2000). A variant of the indicator semivariogram
proportion, and location of different types of sediment bodies approach is transition probability (Markov) models in which the
(e.g., sandstones, shales) from well logs, cores, seismic, or GPR; spatial change from one sediment type (e.g., channel sandstone)
(2) interpret the origin of the sediment bodies; (3) use outcrop to another (e.g., floodplain mudstone) is based on the probability
analogs to predict more sediment-body characteristics; and (4) of the transition. The probability of spatial transition to a particu-
use stochastic (structure-imitating) models to simulate the allu- lar sediment type depends on the existing sediment type, and this
vial architecture between wells, and the rock properties with dependence is called a Markov property. The matrix of probabili-
sediment bodies such as channel-belt sandstones. Stochastic (struc- ties of transition from one sediment type to another can be used
ture-imitating) models are either object-based (also known as to simulate sedimentary sequences in one, two, or three dimen-
discrete or Boolean) or continuous, or both (reviewed by Haldorsen sions (Fig. 59; e.g., Tyler et al., 1994; Doveton, 1994; Carle et al.,
and Damsleth, 1990; Bryant and Flint, 1993; Srivastava, 1994; 1998; Elfeki and Dekking, 2001).
North, 1996; Koltermann and Gorelick, 1996; Dubrule, 1998; and It is commonly difficult to define the input parameters for
Deutsch, 2002). A common combined approach is to use object- stochastic models, especially the semivariograms and transition
based models to simulate the distribution of channel-belt sand- probability matrices in lateral directions. The shapes, dimen-
stone bodies and floodplain shales, and then use continuous sions, and locations of objects in object-based models are difficult
models for simulating “continuous” variables such as porosity to define realistically (see the models of Tyler et al., 1994; Deutsch
and permeability within the objects. and Wang, 1996; Holden et al., 1998; Seifert and Jensen, 1999,
With object-based models, the geometry and orientation of 2000) (Fig. 59). If definition of the dimensions of objects relies
specified objects (e.g., channel-belt sandstone bodies or discrete upon use of outcrop analogs and determination of paleochannel
shales) are determined by Monte Carlo sampling from empirical patterns from subsurface data, there may be serious problems
distribution functions derived mainly from outcrop analogs. (Bridge and Tye, 2000). Process-based models and sequence-
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 145

cal information; (3) detailed understanding of the origin of the


subsurface strata is not necessary in order to use stochastic
models, even though it is desirable; and (4) numerical forward
(process-imitating) models are considered difficult to fit to
subsurface data, and the models and software are not well
developed. Therefore, process-imitating models have had lim-
ited application in quantitative simulation of the architecture of
hydrocarbon reservoirs or aquifers. However, process-imitat-
ing models provide genetic interpretations of deposits and can
predict more realistic sedimentary architecture than structure-
imitating stochastic models. Karssenberg et al. (2001) have
demonstrated that fitting of process-based models to well data
using an essentially trial-and-error approach is possible in
principle. Such an approach involves multiple runs of a process-
based model under different input conditions, and optimiza-
tion of the fitting of output data to observed data. Process-based
models are being developed by Karssenberg et al. (2003), in-
cluding development of software so that models can be fitted to
subsurface data (inversion approach). Another approach is to
use output from process-imitating forward models to provide
input for stochastic models that can be more easily conditioned
with subsurface data.

Process-Based Models of Long-Term,


Large-Scale Erosion in Rivers and Floodplains
Deposition in alluvial valleys is commonly punctuated by
long periods of widespread erosion, resulting in the formation of
incised valleys and river terraces. Long-term, large-scale erosion
in alluvial valleys results from increasing sediment transport rate
in the down-flow direction, such as caused by basin uplift or base-
level fall, or by climatically influenced decrease in upstream
sediment supply. Sediment routing models have not been widely
applied to long-term, large-scale erosion. However, erosion of
fault scarps, valley slopes, and river channels has been modeled
using the diffusion approach (reviewed by Bridge, 2003, p. 339;
Bridge, in press). The diffusion approach can be used to model the
time variation of channel and floodplain erosion arising from an
imposed increase in downstream channel slope. This might occur
as a result of base-level lowering, or local tectonic movements, or
as an avulsing channel flows over the steep edge of an alluvial
ridge. The point in the river profile where the increased slope and
erosion start is called the knickpoint (Fig. 60). As river erosion
proceeds, the knickpoint moves progressively up valley. Incising
channels can still migrate laterally and form floodplains, but
these surfaces would lie beneath and within older, abandoned
floodplains. River terraces are parts of floodplains that have
become elevated above the bankfull level of the active channel as
a result of widespread channel incision. Terrace risers would also
FIG. 58.—Example of 2-D object-based stochastic modeling of experience erosional retreat associated with mass wasting (creep,
alluvial architecture (modified from Srivastava, 1994). debris flows), overland flow, and gullying. As a wave of channel
incision moves upstream, the downstream parts of the valley
experience channel erosion and degradation of terrace risers for
stratigraphic models demonstrate that the spatial distribution of the greatest period of time. Therefore, the floodplain of the
channel-belt sandstones is not random. Unrealistic shapes, di- incising channel should decrease in width in the up-valley direc-
mensions, and spatial distributions of sediment types means that tion (Fig. 60).
it is difficult to get the model to fit observed data and predict Different episodes of degradation and aggradation can result
reservoir or aquifer behavior. Furthermore, because stochastic in a series of terraces of different height and valley fills with a
models do not simulate processes of deposition, they cannot give complicated internal structure (Fig. 61), and the sequence of
any insight into the origin of the alluvial architecture, and they degradation and aggradation may be very difficult to discern
have no predictive value outside the data region. from the pattern of river terraces. In order to reconstruct the
Some reasons why stochastic models are so widely used, timing and location of degradation and aggradation in valley
despite all of the associated problems, are: (1) commercial fills, it is necessary to establish the relative ages of the terrace
software is available; (2) simulations can easily be conditioned surfaces and the deposits beneath them. The relative ages of
using well data, cores, seismic, GPR, and other types of geologi- terrace surfaces have been estimated using the degree of weath-
146 JOHN S. BRIDGE

FIG. 59.—3-D stochastic model simulations. A) Alluvial-fan deposits simulated using a Markov model (Carle et al., 1998). Note the
unrealistic depiction of channel deposits. B) Fluvial deposits simulated using an object model for channel belts and a sequential
indicator simulator for the “background” sheetflood and lacustrine deposits (Seifert and Jensen, 2000). Note the unrealistic
distribution of channel-belt orientations.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 147

A be associated with major earthquakes. Prolonged movement on


alluvial ridge faults or growth of folds over periods of 104 to 105 years could
therefore result in vertical or horizontal motions on the order of
tens to hundreds of meters. Tectonic uplift and subsidence also
knickpoint
vary spatially. For example, the fault zone at the margin of a
floodplain basin comprises many different fault segments that may be
active at different times, and each fault segment has spatially
variable throw along its length. Thus uplift of a mountain belt or
terrace subsidence of a basin comprises the sum of motions along all of
the different active structures. Episodic ground motions of the
order of a meter along a single fault that may be tens of kilome-
incised ters long may locally influence the channel pattern and course
channel belt of a river, but it is unlikely to have a big impact on an entire
deposition
sedimentary basin. However, a succession of movements on
B knickpoint many related faults over, say, 105 years may produce tens to
original bankfull level
hundreds of meters of ground motion affecting basin lengths of
erosion
hundreds of kilometers. Such movements could be responsible
for diversion of major rivers and growth or dissipation of
base-level fall
channel belt alluvial fans (Gawthorpe and Leeder, 2000). If such tectonic
subsequent bankfull level motions were cyclic, they could explain cyclicity in sedimentary
facies over vertical thicknesses of tens to hundreds of meters
and lateral extents of up to hundreds of kilometers.
original floodplain profile

mass wasting channel Evidence of Tectonic Activity in Sedimentary Basins.—


C water erosion erosion

Evidence for tectonic activity in sedimentary basins includes


incised channel belt subsequent floodplain profile variation in: (1) sediment thickness and type in relation to
synsedimentary folds and faults; (2) paleocurrent direction; and
FIG. 60.—Conceptual model for channel-belt incision, knickpoint
and riverbank retreat, and terrace formation resulting from
fall in base level (from Bridge, 2003).

ering of the surface deposits, and using the degree of degradation


of terrace scarps (using diffusion models of scarp retreat). Match-
ing of heights of paired terraces on either side of valleys has also
been used to establish age equivalence. However, terraces may
not be paired on either side of valleys because of the variable age
and elevation of a given abandoned floodplain, or because a once-
present floodplain was later eroded or obscured by colluvium
from valley walls, or because of tectonic deformation of the
terrace. Therefore, it should not be expected that a particular
terrace should have an age and elevation that is constant and
easily distinguishable from others.

Effects of Tectonics on Alluvial Architecture


Tectonic activity controls the rate, amount, and location of
uplift and subsidence, gradient of the land surface, and the
position of rivers and coastlines. Thus, tectonic activity is also
intimately related to climate, vegetation, and eustasy. These
factors in turn control locations, rates, and types of sediment
production, erosion, sediment transport, and deposition. The
effects of tectonic activity on valley slope, channel pattern, and
channel avulsion were discussed above (and reviewed by Schumm
et al., 2000). Thick accumulations of fluvial sediments (sedimen-
tary basins) occur in a variety of different tectonic settings (e.g.,
compressional, extensional, and strike-slip basins), and alluvial
architecture varies with tectonic setting (Miall, 1996).
Tectonic uplift and subsidence can be gradual or episodic
(Schumm et al., 2000). Rates of uplift and subsidence averaged FIG. 61.—Evolution of river terraces and valley fills during falling
over many thousands of years are commonly on the order of and rising base level, based on Gulf of Mexico coastal plain
10-1 mm/year (ranging from 10-2 to 101 mm/year; Leeder, 1999). (from Blum and Price, 1998). Multiple episodes of erosion and
Instantaneous translation on single faults of the order of meters deposition are caused by climate-related variation in supply
typically occurs at intervals of the order of 103 years, and would of water and sediment and in channel avulsion.
148 JOHN S. BRIDGE

(3) sediment provenance. The sediment fills of extensional and sediment routing models are potentially more realistic than
compressional basins are closely related to uplift of adjacent diffusion-based fluvial models, but they require development
uplands due to translation along basin-marginal faults, as dis- before they are capable of simulating alluvial architecture.
cussed below. However, variations in sediment thickness and
type in sedimentary basins can also be related to changes in base Alluvial Architecture in Compressional Basins.—
level or climate in the drainage basin, although such changes
may occur over different spatial and temporal scales. For ex- Examples of compressional basins containing fluvial depos-
ample, the increase in sediment supply to rivers resulting from its are the Paleozoic basins of the Catskill clastic wedge, devel-
tectonically induced headward erosion may occur at a much oped along the western side of the Appalachians in eastern USA
slower rate than the increase in sediment supply due to increase and Canada, the Mesozoic to Tertiary basins on the eastern side
in precipitation in the drainage basin. This is because upstream of the Rockies in North America and the Andes in South America,
migration of a wave of river erosion (a knickpoint) is a local the Tertiary Siwaliks of the Indo-Gangetic basins on the south
process that does not affect all of the drainage basin simulta- edge of the Himalayas, and the Molasse of the European Alps.
neously. In contrast, an increase in precipitation may cause an There are various kinds of compressional basins (e.g., foreland,
increase in erosion of the whole drainage basin simultaneously. foredeep, retroarc; Allen and Allen, 1990; Miall, 1996). They are
Change in mean paleocurrent direction in river deposits may formed by crustal thickening arising from compression, thrust-
indicate tectonically induced change in slope direction or river ing, and folding. The area adjacent to the thickened crust (the
diversion. However, change in paleocurrent direction can also basin) subsides because of gravitational loading. An important
be caused by deposition-induced change in slope direction, and aspect of the crustal flexure is a flexural bulge at the periphery
tectonic activity does not always cause river directions to change of the basin. If the crust has high flexural rigidity and viscosity,
dramatically. Change in sediment provenance may be related to the basin is relatively shallow and wide; otherwise it is deep and
unroofing of new rock types as a result of uplift, or a river narrow. Thus, temporal changes in the rheological properties of
diversion in the hinterland. But erosional exposure of new rock the crust result in changes in the shape of the basin.
types does not require tectonic uplift, and river diversions in the An episode of thrusting, crustal thickening, and loading
hinterland do not need to result in changed provenance of results in uplift and increases in valley slopes, erosion rate, and
sediment supplied to the river. Tectonic activity is also com- sediment supply in the vicinity of the uplift. It also results in
monly associated with volcanic eruptions, and all of the associ- subsidence and deposition in the basin, and growth and migra-
ated features such as catastrophic floods and mudflows, and the tion of the peripheral bulge towards the basin. Erosion of
damming of rivers with debris flows, ash, and lava. uplifted crust and deposition in the foreland basin causes fur-
ther isostatic uplift of the highlands and subsidence in the basin.
General Models for Tectonic Subsidence, The erosional and depositional response to uplift and subsid-
Fluvial Sediment Supply, and Deposition.— ence depend upon the relative timing, positions, and rates of
these events, and these are very difficult to ascertain. For ex-
Paola et al. (1992) developed a theoretical model of large- ample, the nature of crustal subsidence in response to loading
scale variations in mean grain size in alluvial basins, as viewed depends on crustal rheology, specifically whether the crust is
in a section parallel to direction of sediment transport. Sedi- elastic or viscoelastic, and how rheological properties change in
ment transport was modeled using a linear diffusion ap- time with temperature and pressure. The response of weather-
proach, with the diffusivity of sediment being controlled mainly ing, erosion, and sediment supply to changes in source-rock
by water discharge and channel pattern (braided or single type, elevation, slope, and vegetation are difficult to predict.
channel). Grain-size partitioning in the model was based on the The early quantitative models for compressional basins con-
assumption that gravel dominates a deposit until all gravel in sidered flexural isostatic response to loads created by tectonic
transport is exhausted, at which point deposition of sand be- thrusting and sediment deposition, as represented in 2-D sections
gins. They examined the response of an alluvial basin to sinusoi- parallel to the direction of thrusting (e.g., Beaumont, 1981; Jor-
dal variation of four extrinsic controlling variables: rate of dan, 1981; Quinlan and Beaumont, 1984; Beaumont et al., 1988).
sediment supply, diffusivity of sediment, tectonic subsidence Erosion and deposition were not modeled explicitly. Erosion and
rate, and proportion of gravel in the sediment supply (Fig. 62). deposition were modeled explicitly by Flemings and Jordan
The basin response depends strongly on the time scale over (1989) (see also Flemings and Jordan, 1990; Jordan and Flemings,
which variation in the controlling variables occurs. “Slow” and 1990, 1991) using a linear diffusion approach. Paola (2000) com-
“rapid” variations are defined as those that vary with periods pared the stratigraphy predicted by a range of such simple 2-D,
that are respectively longer or shorter than a so-called basin diffusion-based models.
equilibrium time, defined as the square of basin length divided A major step forward in the modeling of deposition at the
by sediment diffusivity. Changes in the rates of uplift of up- river system and basin scale was the treatment of surface pro-
lands, erosion, sediment supply, and subsidence are not linked cesses and drainage-basin evolution in 3-D, and linking them to
in this model, as they must be in nature. Despite the simplifica- climate, tectonic activity, and base-level change (Fig. 63; e.g.,
tions in this model, Heller and Paola (1992) applied it to three Beaumont et al., 1992; Kooi and Beaumont, 1994, 1996; Johnson
alluvial basins to help determine whether conglomerate progra- and Beaumont, 1995; Coulthard et al., 2002; Allen and Densmore,
dation was coincident with tectonic uplift and increase in ero- 2000; Garcia-Castellanos, 2002; Tucker et al., 2002; Clevis et al.,
sion and sediment supply (i.e., syntectonic) or not (antitectonic). 2003). These 3-D basin models indicate that the availability of
Paola et al. (1999) and Marr et al. (2000) further developed this sediment, related to weathering rate and bedrock erodibility,
approach. exerts a strong control on sediment transport rate and basin
Slingerland et al. (1994) coupled a simple sediment routing deposition. Changes in sediment transport rate to basins lag
model with models for tectonic subsidence and uplift, varying behind episodes of tectonic uplift, because of limits to sediment
sea level, and variable sediment supply related to climate change, availability (weathering and bedrock erodibility) and the time it
in order to explore their effects on long profiles of rivers (hence takes for sediment to move downslope through the drainage
long-term, large-scale erosion and deposition). These simple network. Sediment may be stored temporarily in an orogen
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 149

SLOW VARIATION FAST VARIATION


Sediment flux

sand
gravel

basement

Subsidence

Gravel fraction

Diffusivity

FIG. 62.—Hypothetical variation in distribution of sand and gravel in a sedimentary basin subjected to periodic variation in sediment
flux, subsidence, gravel fraction, and sediment diffusivity (from Paola et al., 1992). Figures on left are for slow variation, and lines
are isochrons drawn at 1-million-year intervals. Figures on right are for fast variation and isochrons are drawn every 10,000 years.

because of the development of intermontane basins related to thrusting and uplift, such that relatively coarse sediment fills the
local thrusting and folding (e.g., Tucker and Slingerland, 1996). basin and progrades basinwards (i.e., “syntectonic”), possibly
The lag time may be on the order of 104 to 106 years. resulting in marine regression. Subsequently, subsidence rate
Episodic uplift and subsidence leads to episodic progradation exceeds deposition rate, the coarsest sediment is limited to areas
and retrogradation of fluvial gravels and coastlines, but the near the uplift, and marine transgression may occur. The predic-
relative timing of these events is equivocal (review in Bridge, tions of these models can be changed dramatically by different
2003). In models that assume an elastic crust, tectonic subsidence assumptions about the response to uplift of erosion and transport
in the basin is an immediate response to thrusting, crustal thick- of sediment.
ening, and loading. Thus, subsidence rate may exceed sediment Uplift is likely to be associated with climate change in the
supply and basin deposition rate during crustal loading. Rela- mountain belt and surrounding basins (review in Bridge, 2003, p.
tively coarse sediment produced as a result of uplift is deposited 352). Climate changes are strongly linked to basin stratigraphy,
close to the source, and marine transgression may occur. Such mainly because of the strong link between rainfall, water dis-
basins have been called “underfilled”. Subsequent to uplift (time charge, bedrock erosion rate, and sediment transport rate in
lag on the order of 104 to 106 years), the rate of sediment supply rivers. Climate also indirectly affects basin stratigraphy through
may begin to exceed subsidence rate, and as deposition proceeds its effect on vegetation and weathering rate, which influence both
there is progradation of relatively coarse sediment across the effective precipitation and sediment production. Changes in
basin, and possibly marine regression. Such basins have been sediment transport rate to basins may lag behind changes in
called “overfilled”. Prograding coarse material is called rainfall by on the order of hundreds to thousands of years.
“antitectonic” in this case because it is not coincident with the Burbank (1992) suggested that periods of accelerated isostatic
tectonic uplift. In models that assume viscoelastic crust, sediment uplift associated with climatically induced increase in erosion
supply and deposition rate may exceed subsidence rate during rate should result in deposits that do not vary greatly in thickness
150 JOHN S. BRIDGE

across the basin, because uplift is not associated with downwarp- channel-belt avulsion and local tectonics. Therefore, these mod-
ing of the basin. Eustatic sea-level fluctuations add more com- els do not predict alluvial architecture. However, it is possible to
plexity to basin stratigraphy, especially in near-coastal fluvial link changes in alluvial architecture qualitatively with changes in
deposits (see below). subsidence rate, deposition rate, and grain size. High deposition
The spatial and temporal scales considered in these basin rate of relatively coarse sediment in basins that are back-tilting
models preclude consideration of individual channels and flood- should result in relatively high avulsion frequency close to the
plains, and important controls on alluvial architecture such as uplands, producing fans with nodal avulsion, as observed in the

FIG. 63.—A) 3-D compressional basin model with surface processes (Johnson and Beaumont, 1995).
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 151

A Q

600 kyr 1200 kyr

Q A

800 kyr 1400 kyr


progradation of the coastline,
connection of the gravels

A Q

1000 kyr 1600 kyr


retreat of the coastline and gravel front,
change to lobate gravel distribution

FIG. 63.—B) 3-D compressional-basin model with surface processes and grain-size sorting (Clevis et al., 2003). Perspective views of
successive stages of landscape subjected to tectonic pulsations. Phases of tectonic activity (A) reflected by retreat of coastline and
gravel front. Quiescent periods (Q) associated with progradation of coastline and gravel front.

Himalayan foreland (e.g., the Kosi fan). Overlapping channel sizes of different rivers flowing from the mountain belt. Figure 64
belts on such fans result in sandstone–conglomerate bodies with illustrates diversion of a river by a thrust-related anticline near
large width/thickness ratios, with channel-deposit proportion the edge of a compressional basin, resulting in reduction in the
and connectedness increasing towards the mountain belt (Mackey supply of water and sediment to a basin-margin fan. The size and
and Bridge, 1995). Paleoslopes and river courses are approxi- slope of this fan may then become more influenced by tectonic
mately normal to the edge of the mountain belt (Fig. 64). Periods subsidence than by sediment progradation. The river that re-
of relatively low deposition rate of relatively finer sediment ceives the diverted flow experiences an increased discharge of
should be associated with relatively low avulsion frequency, and water and sediment. Its basin-marginal fan would experience an
the possibility of a relative rise in base level may lead to the increase in deposition rate, and the size and slope of the fan would
highest avulsion frequencies distant from the edge of the moun- become dominated by this sediment progradation. According to
tain belt. In this case, rivers may be flowing parallel or oblique to the sediment routing model used by Robinson and Slingerland
the axis of the foreland basin (Fig. 64). Garcia-Castellanos (2002) (1998a, 1998b), downstream fining of river bed material would
discusses the tectonic influences on the orientation of rivers not be as effective on the growing fan as it would be on the
flowing across foreland basins. shrinking fan. Increase in the discharges and sizes of rivers, and
Tectonic uplift occurs at different rates at different times in in deposition rate and avulsion frequency, on the growing fan
different parts of a mountain belt, which may result in diversions would probably result in increasing channel-deposit proportion
of rivers within mountain belts (e.g., Tucker and Slingerland, and connectedness. Such increases in deposition rate in a cross
1996; Gupta, 1997). Thus, the supply of sediment and water to section oriented normal to the thrust belt and basin axis may not
rivers entering the basin, and the positions of the entry points of be related to a change in uplift rate or subsidence rate in the same
rivers, may vary in space as well as time. Variations in water and cross section. Thus, although tectonism (with or without climate
sediment supply in different rivers may be congruent or incon- change) may occur over a broad region over a long period of time,
gruent. Furthermore, changes in climate in different parts of the the depositional responses may not be the same in different parts
mountain belt (especially during glaciations) could also result in of the basin. This illustrates the desirability of modeling tectonics
both congruent and incongruent changes in the discharges and and sedimentation in three dimensions rather than two.
152 JOHN S. BRIDGE

section c

section d

section c

section d

a
b

300
300

250
250

200
200
thickness (m)

thickness (m)

150
150

100
100

50
50

c d
5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40
distance (km) distance (km)
1800

300 1600

1400
250
thickness (m)

1200
time (kyr)

200
1000

800
150

600
100
400

50 200
e f
5 10 15 20 25 30 35 40 5 10 15 20 25 30 35 40
distance (km) distance (km)
Figure 2.11

FIG. 63 (continued).—C) 3-D compressional-basin model with surface processes and grain-size sorting (Clevis et al., 2003). Cross
sections (c, d) showing distribution of gravel during tectonic pulsation (200,000 yr period). Gravel progrades during tectonic
quiescence. Sinusoidal sea-level fluctuation (period 100,000 years, amplitude 20 m) superimposed on tectonic pulsation shown
in Part E.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 153

FIG. 64.—Hypothetical changes in alluvial-fan size in a foreland (compressional) basin due to hinterland tectonics and river capture
(from Bridge, 2003). Increase in supply of water and sediment to growing fan should result in increased channel size and frequent
avulsion, hence increased channel-deposit proportion.

Compressional basins normally show an increase in deposi- gradient and causes decreasing density and thermal expansion of
tion rate and mean grain size with time. If it is assumed that the lithosphere. This results in both isostatic and expansional
deposition rate is approximately equal to subsidence rate, this uplift at the margins of the thinned lithosphere. Sediment is
trend can be interpreted as an increase in subsidence rate as the eroded from these peripheral uplifts (and other local uplifts) and
thrust front encroaches and/or as the rate of uplift increases. deposited in the extensional basin. As the lithosphere cools,
Inflections in such curves may be related to individual thrusting density increases and subsidence occurs as a result of isostasy and
events. The only way to test the various basin-filling models that thermal contraction. The rate of thermal subsidence decreases as
relate uplift, subsidence, and deposition is to have independent the square root of time. Loading of sediment deposited in the
estimates of the timing and magnitude of uplift and deposition, basin causes downward flexure of the lithosphere and onlap of
which are not normally available. If the thrust front moves sediment at the basin margins. Because flexural rigidity increases
relative to the compressional basin at the same speed as litho- as the lithosphere cools, the zone of onlap increases in width with
spheric plates (order of 10 mm/year), and subsidence rate is on time, and the basin becomes wider and shallower. Thus, lapping
the order of 1 mm/year, then 100 km of crustal convergence of marine sediments onto fluvial sediments is not necessarily due
would produce 10 km of sediment thickness in 10 million years. to eustatic sea-level rise.
Shorter-term episodes of tectonic uplift and subsidence in com- A feature of extensional basins that distinguishes them from
pressional basins possibly range from on the order of a million compressional basins is that the overall subsidence rate decreases
years for major regional thrusting episodes to on the order of a with time. If one side of an extensional basin becomes a “passive”
thousand years for meter-scale throws on single faults. rifted continental margin (e.g., the Atlantic), moves away from
the spreading center for on the order of 100 million years (1000 km
Alluvial Architecture in Extensional Basins.— at 10 mm/year), and average subsidence rate is on the order of 0.1
mm per year, kilometers of sediment could accumulate in the
Examples of extensional basins are the Triassic–Jurassic rift basin. During this time, sea level may be rising eustatically
basins developed during the opening of the Atlantic Ocean, and because of growth of mid-ocean ridges.
now occupying parts of NW Europe, NE North America, Africa, Gawthorpe and Leeder (2000) modeled erosion and deposi-
and South America. Other well-known extensional basins occur tion in extensional basins in relation to the initiation, growth,
in the basin-and-range province of the western United States. propagation, and death of arrays of normal faults (Fig. 65).
Passive continental margins that formed on the Atlantic margins During the fault-initiation stage, rift basins are isolated. Anteced-
following the rifting stage have extensive thickness of Cretaceous ent river courses start to become influenced by the emerging
and Tertiary deposits, much of which are alluvial and deltaic. fault-related topography, and some rivers may be diverted in and
Failed rift basins include the Viking Graben of the North Sea, the along developing rift basins. Incision of new drainage systems in
East Shetland Basin, and the Benue Trough. the uplifted footwalls leads to the development of small, regu-
Extensional basins are caused by lithospheric stretching and larly spaced alluvial fans. The size of the drainage basins and fans
thermal subsidence (reviews in Allen and Allen, 1990; Leeder, decreases towards the fault tips. The centers of these developing
1999). Lithospheric stretching causes brittle fracture and normal rift basins may be occupied by eolian sands, ephemeral or peren-
faulting in relatively shallow parts of the lithosphere but thinning nial lakes, or axial rivers with floodplains (Fig. 65A).
by plastic deformation in deeper parts. Upwelling of hot astheno- Lateral propagation and joining of fault segments leads to
sphere beneath the thinned lithosphere increases the thermal enlargement and coalescence of rift basins and further develop-
154 JOHN S. BRIDGE

Early footwall catchments develop


Shallow lake in easily eroded lithology
Migration and deposition of
aeolian sands controlled by
in hangingwall
depocenter
A
interplay of regional winds and
local structural topography

B A
Early syn-rift depocenter in
growth syncline, bounded
A by fault tip monocline

B B

Lateral propagation and


interaction between fault segments
leads to along-strike onlap Fault scarp with
A incipient drainage
catchments

Aggradational alluvial fans


sourced from incipient
footwall catchments

Fluvial channel belts


preferentially stacked in
centre of early growth syncline
Fluvial plain/playa
A Antecedent drainage incising
Alluvial fans/fan deltas into uplifting footwalls

B Drainage diverted around


Aeolian
propagating fault tips
Lake Pre-rift regional palaeoslope

River channel incision and


terrace formation (high
runoff/low sediment supply) B

Large catchment
at segment boundary B

B
Axial delta sourced Fans incised during
from segment boundary pluvial lake highstand
and axial catchments (high runoff/low
sediment supply)
Shallow highstand
pluvial lake Fault zone with
Aggradational high run off/low
Deep highstand footwall-sourced sediment supply
pluvial lake with fan deltas
basinal turbidites

Fault zone with


Fluvial plain/playa
high run off/high
sediment supply A Antecedent drainage incising
Alluvial fans/fan deltas into uplifting footwalls

Lake Lake deposits B Diversion of antecedent river


through segment boundary
Rift initiation stratigraphy Lake shoreline terraces Abandonded river course

FIG. 65.—Model of alluvial architecture in evolving extensional basins (from Gawthorpe and Leeder, 2000). A) Initiation stage. B)
Interaction and linkage stage.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 155

Large fan 'force' axial river


away from footwall
Increased displacement rate
and low sediment supply lead
to lake development C

A A
B
Alluvial fan toe cut
by axial channel

B
A

B
Tilting and subsidence
Axial progradation of
promotes vertical stacking
fluvio-deltaic system
of axial channel belts adjacent
promoted by development
to footwall fans
of major axial drainage

Tilting and subsidence


of lake bed promotes
vertical stacking of axial
turbidite lobes at base of
Fluvial plain/playa footwall scarp

Alluvial fans/fan deltas A Reversed drainage due to


uplifting footwalls
Lake Lake deposits B Large catchments and fans mark
Rift initiation and interaction breached segment boundaries
stratigraphy Abandoned meander belt

Older footwall-derived fans become


incised as they are uplifted in
footwall of new fault

Main border fault dies and


new fault propagates into
Axial river forced across hangingwall
to hangingwall side of basin
by new footwall fans

Fluvial plain/playa

Alluvial fans/fan deltas

Lake Lake deposits


Rift initiation, interaction and
through-going fault zone
stratigraphy

FIG. 65 (continued).—Model of alluvial architecture in evolving extensional basins (from Gawthorpe and Leeder, 2000).
C) Throughgoing fault stage. D) “Fault death” stage.
156 JOHN S. BRIDGE

ment of drainage systems (Fig. 65B). The largest fans are associ- Seasonal variability of discharge may have an influence on
ated with the larger antecedent rivers and with areas distant from some aspects of channel deposition. In high-latitude and/or
fault tips. Lakes are present in the areas of maximum subsidence, high-altitude rivers, there is commonly a pronounced low-flow
distant from fault tips. Figure 65B illustrates the effects of climate (or frozen) period, and a relatively short spring-to-summer flood
change superimposed on tectonic activity. Increases in precipita- period comprising a number of flood peaks. Furthermore, there
tion and runoff combined with low sediment supply lead to may be marked diurnal variations in flow resulting from daytime
incision of drainage basins, floodplains, and fans, and deposition thawing of ice but night-time freezing. Rivers in seasonal temper-
downstream, possibly in expanded lakes. Such incision does not ate and tropical climates may also have relatively long low-flow
occur if increase in precipitation is accompanied by increase in periods and single-peak or multi-peak flood periods in the spring
sediment supply. Major changes in climate affecting catchments and summer. Seasonal variation in discharge is likely to be
tend to be reflected relatively rapidly (order of thousands of reflected in variation in grain size and sedimentary structures.
years) in downstream alluvial deposition. Episodic uplift of However, it is unlikely that there would be discernible differ-
footwalls causes a wave of incision to move up the catchment, ences in the deposits of rivers in these different climates, insofar
providing increased sediment supply to alluvial fans. However, as they all have low-flow periods and multi-peak flood periods
it may take tens of thousands of years for this to be reflected in when bankfull level would be approached. However, diurnal
increased fan growth, if at all (Allen and Densmore, 2000). Inci- variations in flow in cold climates might be discernible in the
sion of fan channels and local growth of fans can also be associ- character of the stratification (Smith, 1974). The level of base flow
ated with more or less random increases in sediment supply and in perennial rivers can be recognized using features of subaerial
river avulsions. Episodes of fan progradation into the basin result exposure, such as desiccation cracks, footprints, and plant roots.
in upward-coarsening sequences on the order of tens of meters In fact, ephemeral rivers (defined as having a completely dry
thick, whereas fan recession results in fining-upward sequences. course for some of the time) in dry climates can be recognized by
Continued linkage of adjacent fault segments results in defi- observing features of subaerial exposure at the bottom of the
nition of elongate half-graben basins (Fig. 65C), allowing devel- channel. Many published sedimentological criteria (e.g., North
opment of major axial rivers and floodplains bounded by basin- and Taylor, 1996) for the recognition of ephemeral channels (e.g.,
edge fans. Axial channel belts tend to move towards the predominance of planar-laminated and low-angle cross-strati-
downthrown side of the basin, resulting in relatively thick chan- fied sandstone, a lack of well-defined channels or channels with
nel deposits with high channel-deposit proportion and connect- high width/depth ratios, downstream decrease in channel size)
edness interbedded with alluvial-fan deposits. Up-tilted sides are are invalid because they are not distinctive. When assessing the
likely to have low channel-deposit proportion and connected- paleoclimatic significance of ephemeral river channels, it is im-
ness, and well-developed soils. Active tectonism ensures fre- portant to make sure that such channels are not overbank chan-
quent avulsions and episodes of high erosion and deposition rate. nels, which are expected to be ephemeral in any climate. Further-
Deposition in the basin may also be affected by periodic dam- more, although a reduction in channel size in a downstream
ming of rivers by earthquake-induced landslides or volcanic direction could be due to loss of discharge as the river flowed
eruptions (Alexander et al., 1994). across an arid land, it could also be due to a distributive channel
The “fault death” stage (Fig. 65D) is associated with relatively system.
low deposition rates and increasing basin areas, and possibly Deposits of sediment gravity flows are commonly taken as
marine transgression. Figure 65D shows progradation of axial being characteristic of ephemeral rivers in arid climates. Actu-
rivers into a lake or the sea, and relatively low deposition rate of ally, the only prerequisite for a sediment gravity flow is an
axial rivers and floodplains that are being shifted away from the appropriate sediment supply, water, and a relatively large
footwall by deposits of reworked fans. Low channel-deposit slope. Sediment gravity flows occur on the steep slopes of
proportion and connectedness are expected from low avulsion alluvial fans in both arid and wet climates. Furthermore, sedi-
frequency and larger floodplain widths, but high values are ment gravity flows commonly occur down the cut banks of
expected if deposition rate is very low, allowing extensive re- channels, and can be preserved if the channel is being aban-
working of the deposits. High avulsion frequency is expected in doned and filled.
local areas affected by sea-level rise. Some types of floodplain soil are indicative of paleoclimate in
the depositional area (see above), but soil features are controlled
Effect of Climate Change on Alluvial Architecture by parent materials, stage of formation, deposition rate, and
ground-water level, as well as by climate and vegetation. The
Evidence for Paleoclimatic Conditions in Fluvial Deposits.— stage in development of paleosols has been related indirectly to
climate, by associating well-developed paleosols with climati-
Reviews of paleoclimatic indicators in fluvial deposits are cally controlled periods of low deposition rate or erosion. The
given by Miall (1996) and Blum and Törnqvist (2000). Today, composition of clay minerals is commonly related to degree of
rivers of varying size and channel pattern occur in all climatic chemical weathering, hence climate and vegetation. For example,
zones. The suggestion that vegetation (hence climate, or stage in kaolinite and gibbsite are related to the deep weathering in
geological history) has an important influence on channel pat- humid low latitudes. However, the composition of clays depends
terns (i.e., dense vegetation causes low discharge variability, on parent materials and weathering during their journey from
enhances bank cohesion, and promotes the occurrence of single- their site of origin to their site of burial. Therefore, there are likely
channel sinuous rivers) was dismissed previously. Nevertheless, to be many uncertainties in the paleoclimatic meaning of clay
regional changes through time in channel dimensions and pat- minerals.
tern (arising from changes in water and sediment supply during Coal in floodplain environments is not a climate indicator but
floods) may be related to climatic changes. There are many is an indicator of non-saline wetlands where precipitation ex-
examples of a change from large, braided rivers to progressively ceeds evaporation (McCabe, 1984; McCabe and Parrish, 1992).
smaller meandering rivers associated with the change from gla- Lacustrine deposits interbedded with overbank deposits may
cial to interglacial conditions in the Quaternary (reviewed in yield information about climate and changes in climate. For
Blum and Törnqvist, 2000). example, evaporite minerals indicate saline lakes in arid climates,
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 157

and climatically induced changes in lake level may be reflected in plants and their root systems (Driese and Mora, 2001; Algeo et al.,
changes in the relative abundance of evaporites and of terrig- 2001). Larger and deeper roots led to thicker and better-devel-
enous mud and sand. Eolian deposits on floodplains indicate oped paleosols after the Devonian, and the appearance of specific
wind action upon unvegetated sediment surfaces. Although such types of paleosols such as Histosols, Alfisols, Ultisols, and
deposits may merely reflect wind acting on freshly deposited Spodosols. Increase in the rate of weathering due to the spread of
flood sediments, an association with evaporites may indicate land plants, and enhanced preservation of buried organic matter
more prolonged aridity. Fossil plants and animals may suggest because of its resistance to microbes, are thought to have led to a
paleoclimate, especially by comparison with the habitats of com- major drop in atmospheric CO2 from the Devonian to the early
parable extant species. Carboniferous, resulting in global cooling (Berner, 2001; Algeo et
Glacial deposits such as till and ice-contact stratified drift that al., 2001; Driese and Mora, 2001). Similarly, Berner’s model
are interbedded with fluvial deposits are a clear indication of suggests that spread of angiosperms after the Cretaceous may
paleoclimate. However, continental glaciations also bring changes have at least partly influenced the global cooling of the Later
to rivers and floodplains where there is no evidence of deposition Tertiary.
directly from or next to the ice. These changes associated with Cyclicity in climate is related to variations in the Earth’s orbit
glacial and interglacial periods are discussed below. around the Sun and in the Earth’s own rotation, resulting in cyclic
Regional paleogeographic reconstructions can lead to delin- changes in solar radiation. Such Milankovitch cycles have dis-
eation of the areal extent and relative locations of mountain crete periods (on the order of 104 to 105 years) and amplitudes that
ranges, lowlands, and oceans. The elevations of mountains can be interact in a complex way. These cycles had a major influence on
estimated from some types of tectonic model (e.g., Beaumont et the volume of continental ice (hence sea level) during the Pleis-
al., 1988). Paleoslopes and paleodischarges of rivers can be esti- tocene. There has been a high-amplitude 100,000-year period to
mated from sedimentary data. Latitudes can be reconstructed glacial–interglacial cycles during the middle to late Pleistocene,
from paleomagnetic data, and paleoclimate can be estimated but lower-amplitude, 40,000-year cycles dominated prior to that.
using sedimentological and paleontological data. Wind circula- Evidence for these climatic cycles is found in studies of deposi-
tion patterns may be deduced from eolian deposits. As a result of tional sequences, oxygen isotopes of marine microfossils, and
this, rain-shadow effects can be estimated, as can the likelihood of comparative studies of pollen, spores, and marine microfossils.
monsoonal climates. Global climate models have been applied to More recent work on ice cores and marine microfossils indicates
paleogeographies of the past in order to predict paleoclimate and the existence of cycles of abrupt warming followed by longer-
to assess the paleoclimatic evidence that comes from the rocks term cooling over thousands to tens of thousands of years. These
themselves. For example, global climate models have been used global climatic cycles are apparently related to interactions be-
to analyze the onset of the Asian Monsoon in response to the tween the atmosphere, ice masses, and thermohaline circulation
collision of the Indian subcontinent with Asia, and the uplift of of the oceans (Blum and Törnqvist, 2000). Regional climate changes
the Himalayas and the Tibetan Plateau. Such models also attempt are more complex than global changes, because of the way that
to explain the onset of global cooling in the Tertiary, and the changes in the global atmospheric circulation are manifested in
advent of continental glaciation. different regions with differing proportions and elevations of
land. Some short-period (decades and less) fluctuations in re-
Climate Change.— gional climate are related to sunspot cycles or ocean–atmosphere
interactions such as El Niño.
Long-term climate change has been related to changes in The effects of Quaternary climate change on rivers and their
solar output, planetary orbital geometries, geographic distri- deposits have been studied widely (compilations in Gregory,
bution of continents, oceanic circulation patterns, atmospheric 1983; Gregory et al., 1995; Knox, 1983, 1996; Bull, 1991; Benito et
composition, or any combination of these. The last 600 Ma of al., 1999; Blum and Törnqvist, 2000). Some of the changes in
Earth’s history has featured 108 year swings from “icehouse” rivers and their deposits that are expected over glacial–intergla-
climates (latest Precambrian, Late Ordovician–Early Silurian, cial cycles are summarized below and in Table 3. This summary
Permo-Carboniferous, and Late Neogene–Holocene) to “green- is based on much data, but especially data from the Mississippi
house” climates. Icehouse-to-greenhouse swings are coinci- valley and delta plain, the Gulf of Mexico coastal plain, and the
dent with long-term sea-level changes, changes in the miner- Rhine–Meuse delta. Glacial and early deglacial periods are
alogy of oceanic carbonates (“aragonite seas” vs. “calcite seas”; associated with ice-related crustal loading and drainage diver-
Sandberg, 1983), changes in evaporate mineralogy (Hardie, sions in upland areas, increases in flood discharge, mean grain
1996), and changes in the chemistry of sea water (Lowenstein size, and rate of sediment supply, and low sea level. This results
et al., 2001). in progradation from uplands of relatively coarse sediment and
Modeling and proxy data for pCO2 have suggested that most steepening of valley slopes, increases in channel-belt size, and
(but not all) icehouse-to-greenhouse fluctuations were mirrored changes in channel pattern. High avulsion frequency and high
by variations in the carbon dioxide content of the Earth’s atmo- ratios of channel-belt width to floodplain width in up-valley
sphere (Crowley and Berner, 2001; but see Boucot and Gray, 2001, regions should result in increased channel-deposit proportion
and Veizer et al., 2000, for dissenting views). A long-term carbon and connectedness. Falling sea level causes incision or deposi-
cycle controls the carbon dioxide content of the atmosphere. tion in coastal areas (and changing channel patterns) depending
Berner (1991, 1994) and Berner and Kothavala (2001) modeled on factors such as the slope of the exposed surface, rate of sea-
this carbon cycle as controlled by factors such as the degassing of level fall, and up-valley avulsions. Channel-deposit proportion
carbon dioxide from the mantle, recycling of subducted organic and connectedness in coastal areas are increased by larger
and inorganic carbon at magmatic arcs, burial of carbon as channel-belt size and decreased deposition rate but are de-
organic carbon and as carbonates, and consumption of atmo- creased by low avulsion frequency and high floodplain width.
spheric carbon dioxide by carbonate and silicate weathering. Interglacial periods are associated with isostatic uplift and
Silicate and carbonate weathering are apparently controlled partly revegetation of uplands, reduction in flood discharge, grain
by land vegetation. Coincident with the notable climatic swing at size, and rate of sediment supply, and rising sea level. This
the end of the Paleozoic was the development and spread of land results in erosion and terrace formation in up-valley regions.
158 JOHN S. BRIDGE

TABLE 3.—Responses of Mississippi Valley and delta plain to glacial–interglacial cycles


(following Autin et al., 1991; Saucier, 1994).

Parameter Uplands Alluvial valley Coastal plain


Deglaciation-interglacial (Holocene)

Tectonism Isostatic uplift Subsidence increasing towards coast


Sea level Rising:
Transgression
Water supply Decreasing Decreasing
Sediment supply Decreasing Decreasing
Mean sediment size Intermediate Fine
Channel-belt width and depth High Low
Channel pattern Meandering Low sinuosity
Floodplain width Low Very high
Floodplain slope Intermediate Low
Deposition/erosion rate Erosion Terrace formation upvalley
Deposition rate increasing to
coast
Avulsion frequency Intermediate High
Channel-deposit proportion High Low

Glacial maximum-deglaciation (Pleistocene)


Tectonism Crustal loading Subsidence decreasing towards coast
Sea level Low to rising
Water supply High High
Sediment supply High High
Mean sediment size High Intermediate
Channel-belt width and depth High High
Channel pattern Braided Braided-meandering
Floodplain width High Low
Floodplain slope High Intermediate
Deposition/erosion rate Erosion Max. deposition to minor Valley incision to
erosion deposition
Avulsion frequency High, decreasing Low, increasing
Channel-deposit proportion High High, decreasing

Reduced deposition rates, avulsion rates, and channel sizes in interpreted as due to the growth and decay of continental ice.
mid-valley regions lead to decreased channel-deposit propor- There are many potential problems with such interpretations. It
tion and connectedness. Rising sea level leads to increased is not possible to establish the period of these older depositional
deposition rate and avulsion frequency in coastal areas, as well sequences accurately if they are less than 105 years (because of
as increasing channel-deposit proportion and connectedness. dating limitations); therefore, it is not possible to establish the
However, up-valley shift in avulsion points and abandonment age equivalence of depositional sequences with periods less
of delta lobes leads to high overbank deposition rates and small than 105 years in different parts of the world. The links between
channel belts in coastal areas, decreasing channel-deposit pro- solar radiation, climate, sediment production, river and sedi-
portion and connectedness. ment discharges, and sea level are tenuous and poorly known in
These predictions are complicated by local changes in sedi- the absence of major continental ice sheets. Milankovitch-cycle
ment supply, deposition rate, channel pattern, and avulsion periods are similar to those associated with other mechanisms
frequency associated with local tectonism or climate change, not (e.g., tectonism) that can produce sedimentary cycles (Steel et
necessarily congruent with regional changes. Indeed, Blum and al., 1977; Algeo and Wilkinson, 1988; Fraser and DeCelles, 1992).
Törnqvist (2000) describe examples of allostratigraphic units However, some workers have appealed to the 5-to-1 ratio of the
within valley fills that record widespread episodes of river inci- periods of the different Milankovitch cycles (e.g., 20,000 year
sion and aggradation that are interpreted as due to a combination and 100,000 year cycles) to justify orbital forcing of sedimentary
of interglacial–glacial cycles and shorter-term climate changes. cycles.
Each allostratigraphic unit has an erosional base, is on the order Koltermann and Gorelick (1992) explained changes in allu-
of meters to tens of meters thick, extends across valley for kilome- vial-fan deposits due to climate change using a simple sediment
ters to tens of kilometers, and is capped by a well-developed routing model. Floods were generated using a stochastic simu-
paleosol (e.g., Fig. 62). Finally, predictions for present-day tem- lator, and variations in water discharge and sediment transport
perate regions cannot necessarily be expected to apply to regions during large floods were linked to Quaternary paleoclimate
with different climate and vegetation, because changes in water change. The effects of a basin-bounding fault, compaction, and
and sediment supply may be different. base-level change were also considered in this model. Periods of
Many fluvial depositional sequences from meters to tens of high deposition rate of relatively coarse sediment (fan progra-
meters thick (representing 104 to 105 years) in rocks older than dation) were associated with the periods of wet, cool climate
the Pleistocene have been interpreted in terms of Milankovitch when flood discharge of water and sediment were high. Fine
glacio-eustasy (e.g., Olsen, 1990, 1994; Van Tassell, 1994; Read, sediment was deposited during warmer, drier climate.
1994). Also, 100-m-thick, 1 to 10 million-year cycles have been Transcurrent movement on the basin-margin fault caused the
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 159

fan to move horizontally relative to the feeder stream, such that increased frequency of avulsion. Increases in avulsion frequency
successive progradations of the fan were offset relative to each increase channel-deposit proportion and connectedness, but in-
other, producing a 100-m-thick sequence where the thickness of creases in deposition rate and width of flood plains decrease
successive fan deposits either increased or decreased upwards. them. During falling sea level (marine regression), the exposed
land surface may experience erosion or deposition depending on
Effect of Relative Base-Level Change on Alluvial Architecture the slope of the exposed surfaces (Schumm, 1993; Wescott, 1993;
Wood et al., 1993). Erosion occurs on relatively steep slopes given
Eustatic sea-level changes depend greatly upon tectonics and enough time and flow competence. However, erosion of channels
climate, and are due to: (1) changes in the amount of continental and floodplains is a long-term process, and the short-term re-
ice (sea-level changes of more than 100 m over time scales of 104 sponse may be to increase sinuosity of rivers. Also, upstream
to 106 years); (2) changes in the volume of ocean basins, occurring avulsion may cause channel abandonment before incision is
over periods of 10 to 100 million years; and (3) changes in the complete (Leeder and Stewart, 1996). Incision of channels and
amount of water on the surface of the Earth due to outgassing from floodplains is associated with terrace formation, reduction of
the interior of the Earth throughout geological time. Relative sea- valley width, and up-valley migration of knickpoints. Avulsion
level changes depend also on local tectonism and deposition. In frequency is expected to be low in areas of erosion. Channel-
most cases, the effects of relative and eustatic sea-level changes are deposit proportion and connectedness may increase during sea-
combined; therefore, it is generally difficult to recognize a particu- level fall because of reduced deposition rate or erosion and
lar eustatic sea-level change in deposits worldwide, even if the reduced floodplain width, but may decrease as a result of re-
deposits could be dated accurately enough. Exceptions are where duced avulsion frequency.
major eustatic sea-level changes occur over long periods of time. Most alluvial sequence-stratigraphy models have an ero-
The Exxon research group (e.g., Vail et al., 1977) developed a sional base to the sequence (a so-called Type 1 unconformity)
methodology and terminology for describing and analyzing due to valley incision arising from relative fall in base level (Fig.
relatively large-scale depositional sequences with periods rang- 66; Shanley and McCabe, 1993, 1994; Wright and Marriott, 1993;
ing from 105 to 107 years. Because these sequences were initially Gibling and Bird, 1994; Miall, 1996). The erosional base in these
described from seismic records this methodology was referred to models is overlain by superimposed channel-belt deposits, which
as seismic stratigraphy. Subsequently, this approach was also supposedly accumulated under conditions of low deposition
applied to rock outcrops, cores, and well logs, and became known rate and restricted floodplain width (due to valley incision).
as sequence stratigraphy. The jargon associated with sequence Such deposits comprise the so-called “lowstand systems tract”.
stratigraphy is now pervasive, to such an extent that the word Many workers assume that zones of high channel-deposit pro-
“sequence” has developed a special connotation. The Exxon portion in alluvial deposits represent the basal parts of “se-
group interpreted sequences and associated bounding surfaces quences”, and that the basal erosion surface of the lowest
as due to eustatic sea-level changes. These interpretations have sandstone body represents an incised valley (e.g., Aitken and
been criticized by many individuals (notably Miall, 1986, 1991, Flint, 1995; Flint et al., 1995; Hampson et al., 1997; Hampson et
1996) on the grounds that: (1) the method for determining relative al., 1999; Davies et al., 1999). In many cases, evidence for an
sea-level change from seismic records and from cores is flawed; incised valley is lacking. Criteria for incised valleys include
(2) other interpretations of the sequences are not considered; and (Dalrymple et al., 1994): (1) erosional relief that is greater than
(3) the data used to correlate the sequences and interpreted sea- the thickness of a single channel fill; (2) multiple, vertically
level changes worldwide are not generally available. stacked channel bars within the valley; (3) evidence for ex-
There are now many different (sequence stratigraphic) mod- tended periods of nondeposition (mature paleosols) on inter-
els for the effects of relative sea-level change on deposition rate fluves; and (4) alluvial channel deposits resting erosionally
and alluvial architecture (reviewed in Bridge, 2003). Most of them upon shallow marine sands and muds. Commonly, a large
are qualitative and only 2-D, and do not adequately represent all amount of erosional relief on the base of a single channel deposit
of the controls on alluvial deposition. Miall (1991, 1996) has can be misinterpreted as an incised-valley margin (Salter, 1993;
criticized some of the earlier models of the effects of sea-level Best and Ashworth, 1997). Thick and laterally extensive amal-
change on near-coastal alluvial deposition. His main point is that gamated channel deposits are commonly important oil and gas
a relative fall in sea level is not normally associated with alluvial reservoirs. Therefore, such deposits should be interpreted accu-
aggradation, except for the newly exposed part of the sea bed, and rately when predicting their thickness, lateral extent, and bound-
even then only under special circumstances. Whether or not a ing facies. If zones of high channel-deposit proportion are
river valley is incised or aggraded during sea-level fall depends, incorrectly interpreted as incised-valley fills, their extent nor-
among other things, on the slope of the exposed shelf relative to mal to the valley direction will be underestimated, and their
that of the river valley. In general, effects of sea-level change are extent parallel to the valley will be overestimated.
expected to decrease up-valley and be negligible beyond ap- Fluvial deposition may be occurring up-valley while incision
proximately 300 km from the shoreline in a large river like the is occurring near the coast because of falling sea level (Blum and
Mississippi. Price, 1998; Blum and Törnqvist, 2000; Törnqvist et al., 2000; Van
The effect of eustatic sea-level change on fluvial processes Heijst and Postma, 2001), so that the basal erosion surface of the
near shorelines depends on co-variation in tectonic subsidence or sequence will not be coeval or correlatable inland. An ancient
uplift, and deposition or erosion. However, certain generaliza- example of a lowstand valley fill of this kind is given by Willis
tions can be made about the nature of near-coastal alluvial (1997), who explained it in terms of low rate of sea-level change
deposition associated with eustatic sea-level changes that domi- relative to fluvial sediment supply. Interestingly, major erosional
nate over changes in tectonic uplift or subsidence and deposition surfaces may be associated with climate-related fluvial incision
or erosion. During eustatic sea-level rise (marine transgression), or estuarine channels during rising sea level.
slopes of rivers and floodplains are reduced near shore because According to the sequence-stratigraphy models, the depos-
of backwater effects, and the width of valleys increases because of its above the lowstand systems tract were deposited under
drowning. These changes result in reduced grain size of trans- conditions of relatively high deposition rate on a broad alluvial
ported sediment, deposition, change in channel pattern, and plain, in response to rising relative base level (the so-called
160 JOHN S. BRIDGE

FIG. 66.—2-D, qualitative, fluvial sequence-stratigraphy models according to Shanley and McCabe (1993) (left) and Wright and
Marriott (1993) (right). HST = highstand sequence tract, TST = transgressive sequence tract, LST = lowstand sequence tract.

“transgressive systems tract”). The channel-deposit proportion edness as a function of deposition rate and floodplain width.
and connectedness are relatively low as a result. The association However, it is known that sea-level fall is not always accompa-
of relatively high proportion of overbank mud with rising sea nied by valley incision, and the nature of erosion and deposition
level and marine transgression was recognized early on by in alluvial systems is controlled also by climate and tectonism.
McCave (1968, 1969). Törnqvist (1993, 1994) associated high Furthermore, channel-deposit proportion and connectedness
avulsion frequency and anastomosing rivers with periods of are controlled also by factors such as avulsion frequency and
rapid base-level rise and deposition. According to Gibling and channel-belt size. Therefore, it is unlikely that extant 2-D allu-
Bird (1994) and Heckel et al. (1998), extensive coals occur just vial sequence-stratigraphy models are generally applicable. A
below the “maximum flooding surface”, when sea level is near quantitative, 3-D fluvial sequence-stratigraphy model is long
its highstand. However, others predict coal development imme- overdue.
diately above the incised valley fill, associated with the “initial
flooding surface” (Aitken and Flint, 1995; Flint et al., 1995; ACKNOWLEDGMENTS
Hampson et al., 1997, 1999; Davies et al., 1999). Paleosols in the
“transgressive systems tract” are likely to reflect high ground- Comments on the manuscript by Jim Best, Mike Blum, and
water table (Wright and Marriott, 1993). Deposits associated Rudy Slingerland are greatly appreciated.
with the maximum flooding surface may contain evidence of
marine influence. The “highstand systems tract” is also associ- REFERENCES
ated with relatively low channel-deposit proportion, according
to Shanley and McCabe (1993). However, Wright and Marriott AITKEN, J.F., AND FLINT, S.S., 1995, The application of high-resolution
(1993) predict an increase in channel-deposit proportion here, sequence stratigraphy to fluvial systems: a case study from the Upper
and an increase in soil maturity, both related to reduced depo- Carboniferous Breathitt Group, eastern Kentucky, USA: Sedimentol-
sition rate. Retallack (2001) also predicts increasing soil matu- ogy, v. 42, p. 3–30.
rity as sea level rises. AITKEN, M.J., 1998, An Introduction to Optical Dating: Oxford, U.K.,
The alluvial sequence-stratigraphic models discussed above Oxford University Press, 267 p.
all include a basal incised-valley fill related to sea-level fall, and ALEXANDER, J., BRIDGE, J.S., CHEEL, R.J., AND LECLAIR, S.F., 2001, Bed forms
predict variations in channel-deposit proportion and connect- and associated sedimentary structures formed under supercritical
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 161

water flows over aggrading sand beds: Sedimentology, v. 48, ASHWORTH, P.J., BEST, J.L., AND JONES, M., 2004, Relationship between
p.133–152. sediment supply and avulsion frequency in braided rivers: Geology,
ALEXANDER, J., BRIDGE, J.S., LEEDER, M.R., COLLIER, R.E.LL., AND GAWTHORPE, v. 32, p. 21–24.
R.L., 1994, Holocene meander belt evolution in an active extensional ASHWORTH, P.J., BEST, J.L., PEAKALL, J., AND LORSONG, J.A., 1999, The influ-
basin, southwestern Montana: Journal of Sedimentary Research, v. ence of aggradation rate on braided alluvial architecture: field study
B64, p. 542–559. and physical-scale modelling of the Ashburton River gravels, Canter-
ALEXANDER, J., AND LEEDER, M.R., 1987, Active tectonic control on alluvial bury Plains, New Zealand, in Smith, N.D., and Rogers, J., eds., Fluvial
architecture, in Ethridge, F.G., Flores, R.M., and Harvey, M.D., eds., Sedimentology VI: International Association of Sedimentologists,
Recent Developments in Fluvial Sedimentology: SEPM, Special Pub- Special Publication 28, p. 333–346.
lication 39, p. 243–252. ASHWORTH, P.J., BEST, J.L., RODEN, J.E., BRISTOW, C.S., AND KLAASSEN, G.J.,
ALEXANDER, J., AND LEEDER, M.R., 1990, Geomorphology and surface tilting 2000, Morphological evolution and dynamics of a large, sand braid-
in an active extensional basin, SW Montana, USA: Geological Society bar, Jamuna River, Bangladesh: Sedimentology, v. 47, p. 533–555.
of London, Journal, v. 147, p. 461–467. ASLAN, A., AND BLUM, M.D., 1999, Contrasting styles of Holocene avulsion,
ALGEO, T.J., SCHECKLER, S.E., AND MAYNARD, J.B., 2001, Effects of the Middle Texas Gulf Coastal Plain, USA, in Smith, N.D., and Rogers, J., eds.,
to Late Devonian spread of vascular plants on weathering regimes, Fluvial Sedimentology VI: International Association of Sedimentolo-
marine biotas, and global climate, in Gensel, P.G., and Edwards, D., gists, Special Publication 28, p. 193–209.
eds., Plants Invade the Land: New York, Columbia University Press, AUTIN, W.J., BURNS, S.F., MILLER, R.T., SAUCIER, R.T., AND SNEAD, J.I., 1991,
p. 213–236. Quaternary Geology of the lower Mississippi valley, in Morrison,
ALGEO, T.J., AND WILKINSON, B.H., 1988, Periodicity of mesoscale Phanero- R.B., ed., Quaternary Nonglacial Geology: Conterminous U.S: Geo-
zoic sedimentary cycles and the role of Milankovitch orbital modula- logical Society of America, The Geology of North America, v. K-2, p.
tion: Journal of Geology, v. 96, p. 313–322. 547–582.
ALLEN, J.R.L., 1965, A review of the origin and characteristics of recent BAKER, V.R., KOCHEL, R.C., AND PATTON, P.C., 1988, Flood Geomorphology:
alluvial sediments: Sedimentology, v. 5, p. 89–191. New York, Wiley, 528 p.
ALLEN, J.R.L., 1966, On bedforms and palaeocurrents: Sedimentology, v. BAKER, V.R., AND NUMMEDAL, D., 1978, The Channeled Scabland: Washing-
6, p. 153–190. ton, D.C., NASA, Planetary Geology Program, 186 p.
ALLEN, J.R.L., 1970, Physical Processes of Sedimentation: London, George BEAUMONT, C., 1981, Foreland basins: Geophysical Journal of the Royal
Allen & Unwin, 248 p. Astronomical Society, v. 65, p. 291–329.
ALLEN, J.R.L., 1974, Studies in fluviatile sedimentation: implications of BEAUMONT, C., FULLSACK, P., AND HAMILTON, J., 1992, Erosional control of
pedogenic carbonate units, Lower Old Red Sandstone, Anglo-Welsh active compressional orogens, in McClay, K.R., ed., Thrust Tectonics:
outcrop: Geological Journal, v. 9, p. 181–208. London, Chapman & Hall, p. 1–18.
ALLEN, J.R.L., 1978, Studies in fluviatile sedimentation: an exploratory BEAUMONT, C., QUINLAN, G., AND HAMILTON, J., 1988, Orogeny and stratig-
quantitative model for architecture of avulsion-controlled alluvial raphy: numerical models of the Paleozoic in the eastern interior of
suites: Sedimentary Geology, v. 21, p. 129–147. North America: Tectonics, v. 7, p. 389–416.
ALLEN, J.R.L., 1979, Studies in fluviatile sedimentation: an elementary BEHRENSMEYER, A.K., 1987, Miocene fluvial facies and vertebrate taphonomy
geometrical model for the connectedness of avulsion-related channel in northern Pakistan, in Ethridge, F.G., Flores, R.M., and Harvey,
sand bodies: Sedimentary Geology, v. 24, p. 253–267. M.D., eds., Recent Developments in Fluvial Sedimentology: SEPM,
ALLEN, J.R.L., 1982, Sedimentary Structures; Their Character and Physical Special Publication 39, p. 169–176.
Basis, Volume I: Amsterdam, Elsevier Science Publishers, Develop- BEHRENSMEYER, A.K., AND HOOK, R.W., 1992, Paleoenvironmental contexts
ments in Sedimentology 30, 593 p. and taphonomic modes, in Behrensmeyer, A.K., Damuth, J.D.,
ALLEN, P.A., AND ALLEN, J.R., 1990, Basin Analysis: Oxford, U.K., Blackwell, DiMichele, W.A., Potts, R., Sues, H.-D., and Wing, S.L., eds., Terres-
451 p. trial Ecosystems Through Time: Chicago, University of Chicago
ALLEN, P.A., AND DENSMORE, A.L., 2000, Sediment flux from an uplifting Press, p. 15–136.
fault block: Basin Research, v. 12, p. 367–380. BENITO, G., BAKER, V.R., AND GREGORY, K.J., eds., 1999, Palaeohydrology and
ANDERSON, M.P., 1997, Characterization of geological heterogeneity, in Environmental Change: Chichester, U.K., Wiley, 368 p.
Dagan, G., and Neuman, S.P., eds., Subsurface Flow and Transport; A BENNETT, S.J., AND BRIDGE, J.S., 1995a, The geometry and dynamics of low-
Stochastic Approach: Cambridge, U.K., Cambridge University Press, relief bed forms in heterogeneous sediment in a laboratory channel,
p. 23–43. and their relationship to water flow and sediment transport: Journal
ASHMORE, P.E., 1982, Laboratory modeling of gravel braided stream of Sedimentary Research, v. A65, p. 29–39.
morphology: Earth Surface Processes, v. 7, p. 201–225. BENNETT, S.J., AND BRIDGE, J.S., 1995b, An experimental study of flow,
ASHMORE, P.E., 1988, Bed load transport in braided gravel-bed stream bedload transport and bed topography under conditions of erosion
models: Earth Surface Processes and Landforms, v. 13, p. 677–695. and deposition and comparison with theoretical models: Sedimentol-
ASHMORE, P.E., 1991, How do gravel-bar rivers braid?: Canadian Journal ogy, v. 42, p. 117–146.
of Earth Sciences, v. 28, p. 326–341. BERENDSEN, H.J.A., AND STOUTHAMER, E., 2001, Paleogeographic develop-
ASHMORE, P., 1993, Anabranch confluence kinetics and sedimentation ment of the Rhine–Meuse delta, The Netherlands: Assen, Koninklijke
processes in gravel-braided streams, in Best, J.L., and Bristow, C.S., Van Gorcum, 268 p.
eds., Braided Rivers: Geological Society of London, Special Publica- BERES, M., GREEN, A.G., HUGGENBERGER, P., AND HORSTMEYER, H., 1995,
tion 75, p. 129–146. Mapping the architecture of glaciofluvial sediments with three-di-
ASHMORE, P.E., AND CHURCH, M., 1998, Sediment transport and river mensional georadar: Geology, v. 23, p. 1087–1090.
morphology: a paradigm for study, in Klingeman, P.C., Beschta, R.L., BERES, M., HUGGENBERGER, P., GREEN, A.G., AND HORSTMEYER, H., 1999, Using
Komar, P.D., and Bradley, J.B., eds., Gravel-Bed Rivers in the Environ- two- and three-dimensional georadar methods to characterize
ment: Highlands Ranch, Colorado, Water Resources Publications, p. glaciofluvial architecture: Sedimentary Geology, v. 129, p. 1–24.
115–148. BERNER, R.A., 1991, A model for atmospheric CO2 over Phanerozoic time:
ASHWORTH, P.J., 1996, Mid-channel bar growth and its relationship to local American Journal of Science, v. 291, p. 339–376.
flow strength and direction: Earth Surface Processes and Landforms, BERNER, R.A., 1994, GEOCARB II: A revised model of atmospheric CO2
v. 21, p. 103–123. over Phanerozoic time: American Journal of Science, v. 294, p. 56–91.
162 JOHN S. BRIDGE

BERNER, R.A., 2001, The effect of the rise of land plants on atmospheric BRIDGE, J.S., 2001, Characterization of fluvial hydrocarbon reservoirs and
CO2 during the Paleozoic, in Gensel, P.G., and Edwards, D., eds., aquifers: problems and solutions: AAS Revista, Revista de la Asociación
Plants Invade the Land: New York, Columbia University Press, p. de Sedimentología, v. 8, p. 87–114.
173–178. BRIDGE, J.S., 2003, Rivers and Floodplains: Oxford, U.K., Blackwell, 491 p.
BERNER, R.A., AND KOTHAVALA, Z., 2001, GEOCARB III: A revised model of BRIDGE, J S., in press, Numerical modeling of alluvial deposits: recent
atmospheric CO2 over Phanerozoic time: American Journal of Sci- developments, in de Boer, P.L., Postma, G., Kukla, P., van der Zwan,
ence, v. 301, p. 182–204. K., and Burgess, P., eds., Analogue and Numerical Forward Model-
BEST, J.L., 1996, The fluid dynamics of small-scale alluvial bedforms, in ling of Sedimentary Systems; From Understanding to Prediction:
Carling, P.A., and Dawson, M.R., eds., Advances in Fluvial Dynamics International Association of Sedimentologists, Special Publication.
and Stratigraphy: Chichester, U.K., Wiley, p. 67–125. BRIDGE, J.S., ALEXANDER, J., COLLIER, R.E.L., GAWTHORPE, R.L., AND JARVIS, J.,
BEST, J.L., AND ASHWORTH, P.J., 1997, Scour in large braided rivers and the 1995, Ground penetrating radar and coring used to document the
recognition of sequence stratigraphic boundaries: Nature, v. 387, p. large-scale structure of point-bar deposits in 3-D: Sedimentology, v.
275–277. 42, p. 839–852.
BEST, J.L., ASHWORTH, P., BRISTOW, C., AND RODEN, J., 2003, Three-dimen- BRIDGE, J.S., AND BEST, J.L., 1988, Flow, sediment transport and bedform
sional sedimentary architecture of a large, mid-channel sand braid dynamics over the transition from upper-stage plane beds: Implica-
bar, Jamuna River, Bangladesh: Journal of Sedimentary Research, v. tions for the formation of planar laminae: Sedimentology, v. 35, p.
73, p. 516–530. 753–763.
BEST, J.L., AND BRIDGE, J. S., 1992, The morphology and dynamics of low BRIDGE, J.S., AND BEST, J.L., 1997, Preservation of planar laminae arising
amplitude bedwaves upon upper stage plane beds and the preserva- from low-relief bed waves migrating over aggrading plane beds:
tion of planar laminae: Sedimentology, v. 39, p. 737–752. comparison of experimental data with theory: Sedimentology, v. 44,
BIERKENS, M.F.P., AND WEERTS, H.J.T., 1994, Application of indicator simu- p. 253–262.
lation to modeling the lithological properties of a complex confining BRIDGE, J.S., COLLIER, R.E.L., AND ALEXANDER, J., 1998, Large-scale structure
layer: Geoderma, v. 62, p. 265–284. of Calamus River deposits (Nebraska, USA) revealed using ground
BLODGETT, R.H., AND STANLEY, K.O., 1980, Stratification, bed forms, and penetrating radar: Sedimentology, v. 45, p. 977–986.
discharge relations of the Platte River system, Nebraska: Journal of BRIDGE, J.S., AND DIEMER, J.A., 1983, Quantitative interpretation of an
Sedimentary Petrology, v. 50, p. 139–148. evolving ancient river system: Sedimentology, v. 30, p. 599–623.
BLUCK, B.J., 1971, Sedimentation in the meandering River Endrick: Scot- BRIDGE, J.S., AND GABEL, S.L., 1992, Flow and sediment dynamics in a low
tish Journal of Geology, v. 7, p. 93–138. sinuosity, braided river: Calamus River, Nebraska Sandhills: Sedi-
BLUCK, B.J., 1976, Sedimentation in some Scottish rivers of low sinuosity: mentology, v. 39, p. 125–142.
Royal Society of Edinburgh, Transactions, v. 69, p. 425–456. BRIDGE, J.S., AND GORDON, E.A., 1985, Quantitative reconstructions of
BLUCK, B.J., 1979, Structure of coarse grained braided stream alluvium: ancient river systems in the Oneonta Formation, Catskill Magnafacies,
Royal Society of Edinburgh, Transactions, v. 70, p. 181–221. in Woodrow, D.L., and Sevon, W.D., eds., The Catskill Delta: Geologi-
BLUM, M.D., AND PRICE, D.M., 1998, Quaternary alluvial plain construction cal Society of America, Special Paper 201, p. 163–181.
in response to glacio-eustatic and climatic controls, Texas Gulf Coastal BRIDGE, J.S., AND JARVIS, J., 1982, The dynamics of a river bend: a study in
Plain, in Shanley, K.J., and McCabe, P.J., eds., Relative Role of Eustasy, flow and sedimentary processes: Sedimentology, v. 29, p. 499–541.
Climate and Tectonism on Continental Rocks: SEPM, Special Publica- BRIDGE, J.S., AND LEEDER, M.R., 1979, A simulation model of alluvial
tion 59, p. 31–48. stratigraphy: Sedimentology, v. 26, p. 617–644.
BLUM, M.D., AND TÖRNQVIST, T.E., 2000, Fluvial responses to climate and BRIDGE, J.S., AND LUNT, I.A., 2006, Depositional models of braided rivers, in
sea-level change: a review and look forward: Sedimentology, v. 47 Sambrook Smith, G.H., Best, J.L., Bristow, C.S., and Petts, G., eds.,
(Suppl.1), p. 2–48. Braided Rivers; Process, Deposits, Ecology, and Management: Inter-
BOLLA PITTALUGA, M., REPETTO, R., AND TUBINO, M., 2003, Channel bifurca- national Association of Sedimentologists, Special Publication 36, p.
tions in braided rivers: equilibrium configurations and stability: 11–50.
Water Resources Research, v. 39, p. 10–46. BRIDGE, J.S., AND MACKEY, S.D., 1993a, A revised alluvial stratigraphy
BOUCOT, A.J., AND GRAY, J., 2001, A critique of Phanerozoic climatic models model, in Marzo, M., and Puidefàbregas, C., eds., Alluvial Sedimen-
involving changes in the CO2 content of the atmosphere: Earth- tation: International Association of Sedimentologists, Special Publi-
Science Reviews, v. 56, p. 1–159. cation 17, p. 319–337.
BOWN, T.M., AND KRAUS, M.J., 1987, Integration of channel floodplain suite, BRIDGE, J.S., AND MACKEY, S.D., 1993b, A theoretical study of fluvial
I. Developmental sequence and lateral relations of alluvial paleosols: sandstone body dimensions, in Flint, S.S., and Bryant, I.D., eds.,
Journal of Sedimentary Petrology, v. 57, p. 587–601. Geological Modeling of Hydrocarbon Reservoirs: International As-
BRIDGE, J.S., 1978, Paleohydraulic interpretation using mathematical mod- sociation of Sedimentologists, Special Publication 15, p. 213–236.
els of contemporary flow and sedimentation in meandering channels, BRIDGE, J.S., SMITH, N.D., TRENT, F., GABEL, S.L., AND BERNSTEIN, P., 1986,
in Miall, A.D., ed., Fluvial Sedimentology: Canadian Society of Petro- Sedimentology and morphology of a low-sinuosity river: Calamus
leum Geologists, Memoir 5, p. 723–742. River, Nebraska Sand Hills: Sedimentology, v. 33, p. 851–870.
BRIDGE, J.S., 1984, Large-scale facies sequences in alluvial overbank envi- BRIDGE, J.S., AND TYE, R.S., 2000, Interpreting the dimensions of ancient
ronments: Journal of Sedimentary Petrology, v. 54, p. 583–588. fluvial channel bars, channels, and channel belts from wireline-logs
BRIDGE, J.S., 1985, Paleochannel patterns inferred from alluvial deposits: A and cores: American Association of Petroleum Geologists, Bulletin, v.
critical evaluation: Journal of Sedimentary Petrology, v. 55, p. 579– 84, p. 1205–1228.
589. BRIERLEY, G.J., FERGUSON, R.J., AND WOOLFE, K.J., 1997, What is a fluvial
BRIDGE, J.S., 1993, The interaction between channel geometry, water flow, levee?: Sedimentary Geology, v. 114, p. 1–9.
sediment transport and deposition in braided rivers, in Best, J.L., and BRISTOW, C.S., SKELLY, R.L., AND ETHRIDGE, F.G., 1999, Crevasse splays from
Bristow, C.S., eds., Braided Rivers: Geological Society of London, the rapidly aggrading, sand-bed, braided Niobrara River, Nebraska:
Special Publication 75, p. 3–72. effect of base-level rise: Sedimentology, v. 46, p. 1029–1047.
BRIDGE, J.S., 1997, Thickness of sets of cross strata and planar strata as a BRYANT, I.D., AND FLINT, S., 1993, Quantitative clastic reservoir geologi-
function of formative bed-wave geometry and migration, and aggra- cal modeling: problems and perspectives, in Flint, S., and Bryant,
dation rate: Geology, v. 25, p. 971–974. I.D., eds., Geological Modeling of Hydrocarbon Reservoirs: Inter-
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 163

national Association of Sedimentologists, Special Publication 15, Detailed internal architecture of fluvial channel sandstone deter-
p. 3–20. mined from outcrop and 3-D ground penetrating radar: Example
BRYANT, M., FALK, P., AND PAOLA, C., 1995, Experimental study of avulsion from mid-Cretaceous Ferron Sandstone, east-central Utah: American
frequency and rate of deposition: Geology, v. 23, p. 365–369. Association of Petroleum Geologists, Bulletin, v. 85, p. 1583–1608
BUATOIS, L.A., AND MÁNGANO, M.G., 1995, The paleoenvironmental and COTTER, E., 1978, The evolution of fluvial style, with special reference to the
paleoecological significance of the lacustrine Mermia ichnofacies: an central Appalachian Paleozoic, in Miall, A.D., ed., Fluvial Sedimen-
archetypal subaqueous nonmarine trace fossil assemblage: Ichnos, v. tology: Canadian Society of Petroleum Geologists, Memoir 5, p. 361–
4, p. 151–161. 383.
BUATOIS, L.A., MÁNGANO, M.G., GENISE, J.F., AND TAYLOR, T.N., 1998, The COULTHARD, T.J., MACKLIN, M.G., AND KIRKBY, M.J., 2002, A cellular model
ichnologic record of the continental invertebrate invasion: evolution- of Holocene upland river basin and alluvial fan evolution: Earth
ary trends in environmental expansion, ecospace utilization, and Surface Processes and Landforms, v. 27, p. 269–288.
behavioral complexity: Palaios, v. 13, p. 217–240. CROWLEY, K.D., 1983, Large-scale bed configurations (macroforms), Platte
BULL, W.B., 1991, Geomorphic Responses to Climate Change: New York, River Basin, Colorado and Nebraska: Primary structures and formative
Oxford University Press, 326 p. processes: Geological Society of America, Bulletin, v. 94, p. 117–133.
BURBANK, D.W., 1992, Causes of recent Himalayan uplift deduced from CROWLEY, T.J., AND BERNER, R.A., 2001, CO2 and climate change: Science, v.
deposited patterns in the Ganges basin: Nature, v. 357, p. 680–683. 289, p. 270–277.
BURBANK, D.W., AND ANDERSON, R.S., 2001, Tectonic Geomorphology: DALRYMPLE, R.W., BOYD, R., AND ZAITLIN, B.A., eds., 1994, Incised-Valley
Oxford, U.K., Blackwell, 274 p. Systems: Origin and Sedimentary Sequences: SEPM, Special Publica-
CANT, D.J., 1978, Bedforms and bar types in the South Saskatchewan tion 51, 391 p.
River: Journal of Sedimentary Petrology, v. 48, p. 1321–1330. DAVIES, S., HAMPSON, G., FLINT, S., AND ELLIOTT, T., 1999, Continent-scale
CANT, D.J., AND WALKER, R.G., 1976, Development of a braided fluvial sequence stratigraphy of the Namurian, Upper Carboniferous and its
facies model for the Devonian Battery Point Sandstone, Quebec, applications to reservoir prediction, in Fleet, A.J., and Boldy, S.A.R.,
Canada: Canadian Journal of Earth Sciences, v. 13, p. 102–119. eds., Petroleum Geology of Northwest Europe: Proceedings, 5th Con-
CANT, D.J., AND WALKER, R.G., 1978, Fluvial processes and facies sequences ference: Geological Society of London, p. 757–770.
in the sandy braided South Saskatchewan River, Canada: Sedimen- DEUTSCH, C.V., 2002, Geostatistical Reservoir Modeling: New York, Ox-
tology, v. 25, p. 625–648. ford University Press, 376 p.
CARLE, S.F., LABOLLE, E.M., WEISSMANN, G.S., VAN BROCKLIN, D., AND FOGG, DEUTSCH, C., AND COCKERHAM, P., 1994, Practical considerations in the
G.E., 1998, Conditional simulation of hydrofacies architecture: a application of simulated annealing to stochastic simulation: Math-
transition probability/Markov approach, in Fraser, G.S., and Davis, ematical Geology, v. 26, p. 67–82.
J.M., eds., Hydrogeologic Models of Sedimentary Aquifers: SEPM, DEUTSCH, C.V., AND TRAN, T.T., 2002, FLUVSIM: a program for object-
Concepts in Hydrogeology and Environmental Geology 1, p. 147– based modeling of fluvial depositional systems: Computers & Geo-
170. sciences, v. 28, p. 525–535.
CARLING, P.A., 1999, Subaqueous gravel dunes: Journal of Sedimentary DEUTSCH, C.V., AND WANG, L., 1996, Hierarchical object-based stochastic
Research, v. 69, p. 534–545. modeling of fluvial reservoirs: Mathematical Geology, v. 28, p. 857–
CARLING, P.A., GOLZ, E., ORR, H.G., AND RADECKI-PAWLIK, A., 2000a, The 880.
morphodynamics of fluvial sand dunes in the River Rhine, near DIETRICH, W.E., AND SMITH, J.D., 1983, Influence of the point bar on flow in
Mainz, Germany. I. Sedimentology and morphology: Sedimentol- curved channels: Water Resources Research, v. 19, p. 1173–1192.
ogy, v. 47, p. 227–252. DIETRICH, W.E., AND SMITH, J.D., 1984, Bedload transport in a river mean-
CARLING, P.A., WILLIAMS, J.J., GOLZ, E., AND KELSEY, A.D., 2000b, The der: Water Resources Research, v. 20, p. 1355–1380.
morphodynamics of fluvial sand dunes in the River Rhine, near DINEHART, D.L., 1989, Dune migration in a steep, coarse-bedded stream:
Mainz, Germany. II. Hydrodynamics and sediment transport: Sedi- Water Resources Research, v. 25, p. 911–923.
mentology, v. 47, p. 253–278. DINEHART, D.L., 1992, Evolution of coarse gravel bed forms: field measure-
CARLING, P.A., AND SHVIDCHENKO, A.B., 2002, A consideration of the ments at flood stage: Water Resources Research, v. 28, p. 2667–2689.
dune:antidune transition in fine gravel: Sedimentology, v. 49, p. DOVETON, J.H., 1994, Theory and applications of vertical variability mea-
1269–1282. sures from Markov Chain analysis, in Yarus, J.M., and Chambers,
CHANDLER, J.H., 1999, Effective application of automated digital photo- R.L., eds., Stochastic Modeling and Geostatistics: American Associa-
grammetry for geomorphological research: Earth Surface Processes tion of Petroleum Geologists, Computer Applications in Geology, v.
and Landforms, v. 24, p. 51–63. 3, p. 55–64.
CLEVIS, Q., DE BOER, P., AND WACHTER, M., 2003, Numerical modeling of DRIESE, S.G., AND MORA, C.I., 2001, Diversification of Siluro-Devonian
drainage basin evolution and three-dimensional alluvial fan stratig- plant traces in paleosols and influence on estimates of
raphy: Sedimentary Geology, v. 163, p. 85–110. paleoatmospheric CO2 levels, in Gensel, P.G., and Edwards, D., eds.,
COJAN, I., 1999, Carbonate-rich palaeosols in the Late Cretaceous–Early Plants Invade the Land: New York, Columbia University Press, p.
Palaeogene series of the Provence Basin (France), in Thiry, M., and 237–253.
Simon-Coicon, R., eds., Palaeoweathering, Palaeosurfaces and Re- DRIESE, S.G., MORA, C.I., AND ELICK, J.M., 2000, The paleosol record of
lated Continental Deposits: International Association of Sedimen- increasing plant diversity and depth of rooting and changes in
tologists, Special Publication 27, p. 323–335. atmospheric pCO2 in the Siluro-Devonian, in Gastaldo, R.A., and
COLEMAN, J.M., 1969, Brahmaputra River: Channel processes and sedi- DiMichele, W.A., eds., Phanerozoic Terrestrial Ecosystems, A Short
mentation: Sedimentary Geology, v. 3, p. 129–239. Course: New Haven, Connecticut, The Paleontological Society, Pa-
COLLINSON, J.D., 1970, Bedforms of the Tana River, Norway: Geografiska pers, v. 6, p. 47–61.
Annaler, v. 52A, p. 31–56. DUBRULE, O., 1998, Geostatistics in Petroleum Geology: American Asso-
COLLINSON, J. D., 1996, Alluvial sediments, in Reading, H. G., ed., Sedimen- ciation of Petroleum Geologists, Continuing Education Course Note
tary Environments and Facies, Third Edition: New York, Elsevier, p. Series 38, 52 p.
37–82. DULLER, G.A.T., 1996, Recent developments in luminescence dating of
CORBEANU, R.M., SOEGAARD, K., SZERBIAK, R.B., THURMOND, J.B., MCMECHAN, Quaternary sediments: Progress in Physical Geography, v. 20, p.
G.A., WANG, D., SNELGROVE, S.H., FORSTER, C.B., AND MENITOVE, A., 2001, 127–145.
164 JOHN S. BRIDGE

DURY, G.H., 1976, Discharge prediction, present and former, from channel GARCÍA, M., AND NIÑO, Y., 1993, Dynamics of sediment bars in straight and
dimensions: Journal of Hydrology, v. 30, p. 219–245. meandering channels: experiments on the resonance phenomenon:
DURY, G.H., 1985, Attainable standards of accuracy in the retrodiction of Journal of Hydraulic Research, v. 31, p. 739–761.
palaeodischarge from channel dimensions: Earth Surface Processes GARCIA-CASTELLANOS, D., 2002, Interplay between lithospheric flexure
and Landforms, v. 10, p. 205–213. and river transport in foreland basins: Basin Research, v. 14, p. 89–
EKDALE, A.A., BROMLEY, R.G., AND PEMBERTON, S.G., 1984, Ichnology: Trace 104.
Fossils in Sedimentology and Stratigraphy: SEPM, Short Course 15, GAWTHORPE, R.L., COLLIER, R.E.LL., ALEXANDER, J., BRIDGE, J.S., AND LEEDER,
317 p. M.R., 1993, Ground penetrating radar: application to sandbody ge-
ELFEKI, A., AND DEKKING, M., 2001, A markov chain model for subsurface ometry and heterogeneity studies, in North, C.J., and Prosser, D.J.,
characterization: theory and application: Mathematical Geology, v. eds., Characterization of Fluvial and Aeolian Reservoirs: Geological
33, p. 569–589. Society of London, Special Publication 73, p. 421–432.
ELLIOTT, T., 1974, Interdistributary bay sequences and their genesis: Sedi- GAWTHORPE, R.L., AND LEEDER, M.R., 2000, Tectono-sedimentary evolution
mentology, v. 21, p. 611–622. of active extensional basins: Basin Research, v. 12, p. 195–218.
ENOS, P., 1991, Sedimentary parameters for computer modeling, in GENISE, J.G., MÁNGANO, M.G., BUATOIS, L.A., LAZA, J.H., AND VERDE, M., 2000,
Franseen, E.K., Watney, W.L., Kendall, C.G.St.C., and Ross, W., eds., Insect trace fossil associations in paleosols: the Coprinisphaera ichno-
Sedimentary Modeling: Computer Simulations and Methods for facies: Palaios, v. 15, p. 49–64.
Improved Parameter Definition: Kansas Geological Survey, Bulletin GENSEL, P.G., AND EDWARDS, D., 2001, Plants Invade the Land: New York,
233, p. 63–99. Columbia University Press, 304 p.
ETHRIDGE, F.G., AND SCHUMM, S.A., 1978, Reconstructing paleochannel GERMANOWSKI, D., AND SCHUMM, S.A., 1993, Changes in braided river
morphologic and flow characteristics: methodology, limitations and morphology resulting from aggradation and degradation: Journal of
assessment, in Miall, A.D., ed., Fluvial Sedimentology: Canadian Geology, v. 101, p. 451–466.
Society of Petroleum Geologists, Memoir 5, p. 703–721. GIBLING, M.R., AND BIRD, D.J., 1994, Late Carboniferous cyclothems and
ETHRIDGE, F.G., SKELLY, R.L., AND BRISTOW, C.S., 1999, Avulsion and crevassing alluvial paleovalleys in the Sydney Basin, Nova Scotia: Geological
in the sandy, braided Niobrara River: complex response to base-level Society of America, Bulletin, v. 106, p. 105–117.
rise and aggradation, in Smith, N.D., and Rogers, J., eds., Fluvial GORDON, E.A., AND BRIDGE, J.S., 1987, Evolution of Catskill (Upper Devo-
Sedimentology VI: International Association of Sedimentologists, nian) river systems: Journal of Sedimentary Petrology, v. 57, p. 234–
Special Publication 28, p. 179–191. 249.
FARRELL, K.M., 1987, Sedimentology and facies architecture of overbank GRAN, K., AND PAOLA, C., 2001, Riparian vegetation controls on braided
deposits of the Mississippi River, False River Region, Louisiana, in stream dynamics: Water Resources Research, v. 37, p. 3275–3283.
Ethridge, F.G., Flores, R.M., and Harvey, M.D., eds., Recent Develop- GREGORY, K.J., ed., 1983, Background to Palaeohydrology: Chichester,
ments in Fluvial Sedimentology: SEPM, Special Publication 39, p. U.K., Wiley, 486 p.
111–120. GREGORY, K.J., STARKEL, L., AND BAKER, V.R., eds., 1995, Global Continental
FARRELL, K.M., 2001, Geomorphology, facies architecture, and high reso- Palaeohydrology: Chichester, U.K., Wiley, 346 p.
lution, non-marine sequence stratigraphy in avulsion deposits, GUPTA, S., 1997, Himalayan drainage patterns and the origin of fluvial
Cumberland marshes, Saskatchewan: Sedimentary Geology, v. 139, megafans in the Ganges foreland basin: Geology, v. 25, p. 11–14.
p. 93–150. HALDORSEN, H.H., AND DAMSLETH, E., 1990, Stochastic modeling: Journal of
FASTOVSKY, D.E., AND MCSWEENEY, K., 1987, Paleosols spanning the Creta- Petroleum Technology, v. 42, p. 404–412.
ceous–Paleogene transition, eastern Montana and western North HAMPSON, G.J., ELLIOTT, T., AND DAVIES, S.J., 1997, The application of
Dakota: Geological Society of America, Bulletin, v. 99, p. 66–77. sequence stratigraphy to Upper Carboniferous fluvio-deltaic strata of
FLEMINGS, P.B., AND JORDAN, T.E., 1989, A synthetic stratigraphic model of the onshore UK and Ireland: implications for the southern North Sea:
foreland basin development: Journal of Geophysical Research, v. 94, Geological Society of London, Journal, v. 154, p. 719–733.
p. 3851–3866. HAMPSON, G.J., DAVIES, S.J. ELLIOTT, T., FLINT, S.S., AND STOLLHOFEN, H., 1999,
FLEMINGS, P.B., AND JORDAN, T. E., 1990, Stratigraphic modeling of foreland Incised valley fill sandstone bodies in Upper Carboniferous fluvio–
basins: Interpreting thrust deformation and lithospheric rheology: deltaic strata: recognition and reservoir characterization of Southern
Geology, v. 18, p. 430–434. North Sea analogues, in Fleet, A.J., and Boldy, S.A.R., eds., Petroleum
FLINT, S.S., AITKEN, J., AND HAMPSON, G., 1995, Application of sequence Geology of Northwest Europe: Proceedings 5th Conference: The
stratigraphy to coal-bearing coastal plain successions: implications Geological Society of London, p. 771–788.
for the UK Coal measures, in Whateley, M.K.G., and Spears, D.A., HARBOR, D.J., 1998, Dynamics of bedforms in the lower Mississippi River:
eds., European Coal Geology: Geological Society of London, Special Journal of Sedimentary Research, v. 68, p. 750–762.
Publication 82, p. 1–16. HARDIE, L.A., 1996, Secular variations in seawater chemistry: An explana-
FRASER, G.S., AND DECELLES, P.G., 1992, Geomorphic controls on sediment tion for the coupled secular variation in the mineralogies of marine
accumulation at margins of foreland basins: Basin Research, v. 4, p. limestones and potash evaporates over the past 600 m.y.: Geology, v.
233–252. 24, p. 279–283.
FREDERICI, B., AND PAOLA, C., 2003, Dynamics of channel bifurcations in HASIOTIS, S.T., 2002, Continental Trace Fossils: SEPM, Short Course Notes
noncohesive sediments: Water Resources Research, v. 39, p. 1162. No. 51, 132 p.
FREY, R.W., PEMBERTON, S.G., AND FAGERSTROM, J.A., 1984, Morphologi- HASIOTIS, S.T., 2004, Reconnaissance of Upper Jurassic Morrison Forma-
cal, ethological and environmental significance of the ichnogenera tion ichnofossils, Rocky Mountain Region, USA: paleoenvironmen-
Scoyenia and Ancorichnus: Journal of Paleontology, v. 58, p. 511– tal, stratigraphic, and paleoclimatic significance of terrestrial and
528. freshwater ichnocoenoses: Sedimentary Geology, v. 167, p. 177–268.
FUJITA, I., 1989, Bar and channel formation in braided streams, in Ikeda, S., HASIOTIS, S.T., AND DUBIEL, D.L., 1995, Termite (Insecta, Isoptera) nest
and Parker, G., eds., River Meandering: American Geophysical Union, ichnofossils from the Upper Triassic Chinle Formation, Petrified
Water Resources Monograph 12, p. 417–462. Forest National Monument, Arizona: Ichnos, v. 4, p.111–130.
GABEL, S.L., 1993, Geometry and kinematics of dunes during steady and HECKEL, P.H., GIBLING, M.R., AND KING, N.R., 1998, Stratigraphic model for
unsteady flows in the Calamus River, Nebraska, USA: Sedimentol- glacial-eustatic Pennsylvanian cyclothems in highstand nearshore
ogy, v. 40, p. 237–269. detrital regimes: Journal of Geology, v. 106, p. 373–383.
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 165

HELLER, P.L., AND PAOLA, C., 1992, The large scale dynamics of grain-size KARSSENBERG, D., BRIDGE, J.S., STOUTHAMER, E., KLEINHANS, M.G., AND
variation in alluvial basins, 2. Application to syntectonic conglomer- BERENDSEN, H.J.A., 2003, Modelling cycles of fluvial aggradation and
ate: Basin Research, v. 4, p. 91–102. degradation using a process-based alluvial stratigraphy model. Pro-
HELLER, P.L., AND PAOLA, C., 1996, Downstream changes in alluvial archi- ceedings of “Analogue and Numerical Modelling of Sedimentary
tecture: an exploration of controls on channel-stacking patterns: Systems; From Understanding to Prediction”, 9–11 Oct., 2003, Utre-
Journal of Sedimentary Research, v. B66, p. 297–306 cht, the Netherlands.
HELLER, P.L., PAOLA, C., HWANG, I-G., JOHN, B., AND STEEL, R., 2001, Geomor- KARSSENBERG, D., DALMAN, R., WELTJE, G.J., POSTMA, G., AND BRIDGE, J.S., 2004,
phology and sequence stratigraphy due to slow and rapid base-level Numerical modelling of delta evolution by nesting high and low
changes in an experimental subsiding basin (XES 96-1): American resolution process-based models of sedimentary basin filling: Joint
Association of Petroleum Geologists, Bulletin, v. 85, p. 817–838. EURODELTA/EUROSTRATAFORM meeting. Venice (Italy), Octo-
HOEY, T., AND SUTHERLAND, A.J., 1991, Channel morphology and bedload ber 20–23, 2004, p. 1–37.
pulses in braided rivers: a laboratory study: Earth Surface Processes KHAN, I.A., BRIDGE, J.S., KAPPELMAN, J., AND WILSON, R., 1997, Evolution of
and Landforms, v. 16, p. 447–462. Miocene fluvial environments, eastern Potwar plateau, northern
HOLBROOK, J., AND SCHUMM, S.A., 1999, Geomorphic and sedimentary Pakistan: Sedimentology, v. 44, p. 221–251.
response of rivers to tectonic deformation: a brief review and critique KLEINHANS, M.G., WILBERS, A.W.E., DE SWAAF, A., AND VAN DEN BERG, J.H.,
of a tool for recognizing subtle epeirogenic deformation in modern 2002, Sediment supply–limited bedforms in sand–gravel rivers: Jour-
and ancient settings: Tectonophysics, v. 305, p. 287–306. nal of Sedimentary Research, v. 72, p. 629–640.
HOLDEN, L., HAUGE, R., SKARE, Ø., AND SKORSTAD, A., 1998, Modeling of fluvial KNAAPEN, M.A.F., HULSCHER, S.J.M.H., DE VRIEND, H.J., AND VAN HARTEN, A.,
reservoirs with object models: Mathematical Geology, v. 30, p. 473–496. 2001, Height and wavelength of alternate bars in rivers: modeling and
HUGGENBERGER, P., 1993, Radar facies: recognition of characteristic braided laboratory experiments: Journal of Hydraulic Research, v. 39, p. 147–
river structures of the Pleistocene Rhine gravel ( NE part of Switzer- 153.
land ), in Best, J.L., and Bristow, C.S., eds., Braided Rivers: Geological KNIGHT, D.W., AND BROWN, F.A., 2001, Resistance studies of overbank flow
Society of London, Special Publication 75, p. 163–176. in rivers with sediment using the flood channel facility: Journal of
HUGHES, D.A., AND LEWIN, J., 1982, A small-scale flood-plain: Sedimentol- Hydraulics Research, v. 39, p. 283–302.
ogy, 29, p. 891–895. KNIGHT, D.W., AND SHIONO, K., 1996, River channel and floodplain hydrau-
JACKSON, R.G., 1976, Large-scale ripples of the Lower Wabash River: lics, in Anderson, M.G., Walling, D.E., and Bates, P.D., eds., Flood-
Sedimentology, v. 23, p. 593–623. plain Processes: Chichester, U.K., Wiley, p. 139–182.
JOHNSON, D.D., AND BEAUMONT, C., 1995, Preliminary results from a plan- KNOX, J.C., 1983, Responses of river systems to Holocene climates, in
form kinematic model of orogen evolution, surface processes and the Wright, H.E., and Porter, S.C., eds., Late Quaternary Environments of
development of clastic foreland basin stratigraphy, in Dorobek, S.L., the United States, vol. 2, The Holocene: Minneapolis, University of
and Ross, G.M., eds., Stratigraphic Evolution of Foreland Basins: Minnesota Press, p. 26–41.
SEPM, Special Publication 52, p. 3–24. KNOX, J.C., 1996, Fluvial systems since 20,000 years BP, in Gregory, K.J.,
JOHNSON, N.M., AND DREISS, S.J., 1989, Hydrostratigraphic interpretation Starkel, L., and Baker, V.R., eds., Continental Palaeohydrology:
using indicator geostatistics: Water Resources Research, v. 25, p. Chichester, U.K., Wiley, p. 87–108,
2501–2510. KOCHEL, R.C., AND BAKER, V.R., 1988, Palaeoflood analysis using slack
JOL, H.M., 1995, Ground penetrating radar antennae frequencies and water deposits, in Baker, V.R., Kochel, R.C., and Patton, P.C., eds.,
transmitter powers compared for penetration depth, resolution and Flood Geomorphology: New York, Wiley, p. 357–376.
reflection continuity: Geophysical Prospecting, v. 43, p. 693–709. KOLTERMANN, C.E., AND GORELICK, S.M., 1992, Paleoclimatic signature in
JOL, H.M., AND SMITH, D.G., 1991, Ground penetrating radar of northern terretrial flood deposits: Science, v. 256, p. 1775–1782.
lacustrine deltas: Canadian Journal of Earth Sciences, v. 28, p. 1939– KOLTERMANN, C.E., AND GORELICK, S.M., 1996, Heterogeneity in sedimentary
1947. deposits: a review of structure-imitating, process-imitating and de-
JONES, L.S., AND SCHUMM, S.A., 1999, Causes of avulsion: an overview, in scriptive approaches: Water Resources Research, v. 32, p. 2617–2658.
Smith, N.D., and Rogers, J., eds., Fluvial Sedimentology VI: Interna- KOMAR, P.D., 1996, Entrainment of sediments from deposits of mixed
tional Association of Sedimentologists, Special Publication 28, p. 171– grain sizes and densities, in Carling, P.A., and Dawson, M.R., eds.,
178. Advances in Fluvial Dynamics and Stratigraphy: Chichester, U.K.,
JORDAN, T.E., 1981, Thrust loads and foreland basin evolution, Cretaceous Wiley, p. 127–181.
western United States: American Association of Petroleum Geolo- KOOI, H., AND BEAUMONT, C., 1994, Escarpment evolution on high-elevation
gists, Bulletin, v. 65, p. 2506–2620. rifted margins: insights derived from a surface processes model that
JORDAN, T.E., AND FLEMINGS, P.B., 1990, From geodynamical models to basin combines diffusion, advection, and reaction: Journal of Geophysical
fill—a stratigraphic perspective, in Cross, T.A., ed., Quantitative Research, v. 99, p. 12,191–12,209.
Dynamic Stratigraphy: Englewood Cliffs, New Jersey, Prentice-Hall, KOOI, H., AND BEAUMONT, C., 1996, Large-scale geomorphology: classical
p. 149–163. concepts reconciled and integrated with contemporary ideas using a
JORDAN, T.E., AND FLEMINGS, P.B., 1991, Large-scale stratigraphic architec- surface processes model: Journal of Geophysical Research, v. 101, p.
ture, eustatic variation, and unsteady tectonism: a theoretical evalu- 3361–3386.
ation: Journal of Geophysical Research, v. 96, p. 6681–6699. KOSS, J.E., ETHRIDGE, F.G., AND SCHUMM, S.A., 1994, An experimental study
JULIEN, P.Y., AND KLAASEN, G.J., 1995, Sand-dune geometry of large rivers of the effects of base-level change on fluvial, coastal plain, and shelf
during floods: American Society of Civil Engineers, Journal of Hy- systems: Journal of Sedimentary Research, v. B64, p. 90–98.
draulic Engineering, v. 121, p. 657–663. KRAUS, M.J., 1987, Integration of channel and floodplain suites, II. Vertical
JULIEN, P.Y., KLAASSEN, G.J., TEN BRINKE, W.B.M., AND WILBERS, A.W.E., 2002, relations of alluvial paleosols: Journal of Sedimentary Petrology, v.
Case study: bed resistance of Rhine River during 1998 flood: Ameri- 56, p. 602–612.
can Society of Civil Engineers, Journal of Hydraulic Engineering, v. KRAUS, M.J., 1996, Avulsion deposits in lower Eocene alluvial rocks,
128, p. 1042–1050. Bighorn Basin, Wyoming: Journal of Sedimentary Research, v. B66, p.
KARSSENBERG, D., TÖRNQVIST, T.E., AND BRIDGE, J.S., 2001, Conditioning a 354–363.
process-based model of sedimentary architecture to well data: Jour- KRAUS, M.J., 1999, Paleosols in clastic sedimentary rocks: their geologic
nal of Sedimentary Research, v. 71, p. 868–879. applications: Earth-Science Reviews, v. 47, p. 41–70.
166 JOHN S. BRIDGE

KRAUS, M.J., AND ASLAN, A., 1999, Palaeosol sequences in floodplain LEEDER, M.R., AND STEWART, M., 1996, Fluvial incision and sequence stratig-
environments: a hierarchical approach, in Thiry, M., and Simon- raphy: alluvial responses to relative base level fall and their detection
Coicon, R., eds., Palaeoweathering, Palaeosurfaces and related Con- in the geological record, in Hesselbo, S.P., and Parkinson, D.N., eds.,
tinental Deposits: International Association of Sedimentologists, Spe- Sequence Stratigraphy in British Geology: Geological Society of Lon-
cial Publication 27, p. 303–321. don, Special Publication 103, p. 47–61.
KRAUS, M.J., AND GWINN, B.M., 1997, Controls on the development of early LEVEY, R.A., 1978, Bedform distribution and internal stratification of
Eocene avulsion deposits and floodplain paleosols, Willwood Forma- coarse-grained point bars, Upper Congaree River, South Carolina, in
tion, Bighorn Basin: Sedimentary Geology, v. 114, p. 33–54. Miall, A.D., ed., Fluvial Sedimentology: Canadian Society of Petro-
KRAUS, M.J., AND WELLS, T.M., 1999, Recognizing avulsion deposits in the leum Geologists, Memoir 5, p. 105–127.
ancient stratigraphic record, in Smith, N.D., and Rogers, J., eds., LOWENSTEIN, T.K., TIMOFEEFF, M.N., BRENNAN, S.T., AND DEMICCO, R.V., 2001,
Fluvial Sedimentology VI: International Association of Sedimentolo- Oscillations in Phanerozoic seawater chemistry: Evidence from fluid
gists, Special Publication 28, p. 251–268. inclusions: Science, v. 294, p. 1086–1088.
LAI, C.-J., LIU, C.-L., AND LIN, Y.-Z., 2000, Experiments on flood-wave LUNT, I.A., AND BRIDGE, J.S., 2004, Evolution and deposits of a gravelly braid
propagation in compound channel: American Society of Civil Engi- bar and a channel fill, Sagavanirktok river, Alaska: Sedimentology, v.
neers, Journal of Hydraulic Engineering, v. 126, p. 492–501. 51, p. 415–432.
LANE, S.N., CHANDLER, J.H., AND PORFIRI, K., 2001, Monitoring river channel LUNT, I.A., BRIDGE, J.S., AND TYE, R.S., 2004a, Development of a 3-D
and flume surfaces with digital photogrammetry: American Society depositional model of braided river gravels and sands to improve
of Civil Engineers, Journal of Hydraulic Engineering, v. 127, p. 871– aquifer characterization, in Bridge, J.S., and Hyndman, D., eds.,
877. Aquifer Characterization: SEPM, Special Publication 80, p. 139–
LANE, S.N., CHANDLER, J.H., AND RICHARDS, K.S., 1994, Developments in 169.
monitoring and terrain modeling small-scale river-bed topography: LUNT, I.A., BRIDGE, J.S., AND TYE, R.S., 2004b, A quantitative, three-dimen-
Earth Surface Processes and Landforms, v. 19, p. 349–368. sional depositional model of gravelly braided rivers: Sedimentology,
LANE, S.N., RICHARDS, K.S., AND CHANDLER, J.H., 1995, Morphological v. 51, p. 377–414.
estimation of the time-integrated bedload transport rate: Water Re- MACK, G.H., JAMES, W.C., AND MONGER, H.C., 1993, Classification of paleo-
sources Research, v. 31, p. 761–772. sols: Geological Society of America, Bulletin, v. 105, p. 129–136.
LANE, S.N., RICHARDS, K.S., AND CHANDLER, J.H., EDS., 1998, Landform MACK, G.H., AND LEEDER, M.R., 1998, Channel shifting of the Rio grande,
Monitoring, Modeling and Analysis: Chichester, U.K., Wiley, 466 p. southern Rio Grande rift: implications for alluvial stratigraphy mod-
LANE, S.N., WESTAWAY, R.M., AND HICKS, D.M., 2003, Estimation of erosion els: Sedimentary Geology, v. 177, p. 207–219.
and deposition volumes in a large gravel-bed, braided river using MACKEY, S.D., AND BRIDGE, J.S., 1995, Three dimensional model of alluvial
synoptic remote sensing: Earth Surface Processes and Landforms, v. stratigraphy: theory and application: Journal of Sedimentary Re-
28, p. 249–271. search, v. B65, p. 7–31.
LANZONI, S., 2000a, Experiments on bar formation in straight flume. 1. MAKASKE, B., 2001, Anastomosing rivers: a review of their classification,
uniform sediment: Water Resources Research, v. 36, p. 3337–3349. origin and sedimentary products: Earth-Science Reviews, v. 53, p.
LANZONI, S., 2000b, Experiments on bar formation in straight flume. 2. 149–196.
graded sediment: Water Resources Research, v. 36, p. 3351–3363. MARR, J.G., SWENSON, J.B., PAOLA, C., AND VOLLER, V.R., 2000, A two-
LECLAIR, S.F., 2002, preservation of cross-strata due to the migration of diffusion model of fluvial stratigraphy in closed depositional basins:
subaqueous dunes: an experimental investigation: Sedimentology, v. Basin Research, v. 12, p. 381–398.
49, p. 1157–1180. MARTINI, I.P., BAKER, V.R., AND GARZON, G., EDS., 2002, Flood and Megaflood
LECLAIR, S.F., AND BRIDGE, J.S., 2001, Quantitative interpretation of sedi- Processes and Deposits: International Association of Sedimentolo-
mentary structures formed by river dunes: Journal of Sedimentary gists, Special Publication 32, 312 p.
Research, v. 71, p. 713–716. MCCABE, P.J., 1984, Depositional environments of coal and coal-bearing
LEDDY, J.O., ASHWORTH, P.J., AND BEST, J.L., 1993, Mechanisms of anabranch strata, in Rahmani, R.A., and Flores, R.M., eds., Sedimentology of
avulsion within gravel-bed rivers: observations from a physical scale Coal and Coal-Bearing Sequences: International Association of Sedi-
model, in Best, J.L., and Bristow, C.S., eds., Braided Rivers: Geological mentologists, Special Publication 7, p. 13–42.
Society of London, Special Publication 75, p. 119–127. MCCABE, P.J., AND PARRISH, J.T., 1992, Tectonic and climatic controls on the
LEEDER, M.R., 1975, Pedogenic carbonates and flood sediment accretion distribution and quality of Cretaceous coals, in McCabe, P.J., and
rates: A quantitative method for alluvial arid-zone lithofacies: Geo- Parrish, J.T., eds., Controls on the Distribution and Quality of Creta-
logical Magazine, v. 112, p. 257–270. ceous Coals: Geological Society of America, Special Paper 267, p. 1–15.
LEEDER, M.R., 1978, A quantitative stratigraphic model for alluvium with MCCARTHY, P.J., AND PLINT, A.G., 1998, Recognition of interfluve sequence
special reference to channel deposit density and interconnectedness, boundaries: integrating paleopedology and sequence stratigraphy:
in Miall, A.D., ed., Fluvial Sedimentology: Canadian Society of Petro- Geology, v. 26, p. 387–390.
leum Geologists, Memoir 5, p. 587–596. MCCAVE, I.N., 1968, Shallow and marginal marine sediments associated
LEEDER, M.R., 1993, Tectonic controls upon drainage basin development, with the Catskill complex in the Middle Devonian of New York, in
river channel migration and alluvial architecture: implications for hy- Klein, G.deV., ed., Late Paleozoic and Mesozoic Continental Sedi-
drocarbon reservoir development and characterization, in North, C.P., mentation, Northeastern North America: Geological Society of
and Prosser, D.J., eds., Characterization of Fluvial and Aeolian Reser- America, Special Paper 106, p. 75–108.
voirs: Geological Society of London, Special Publication 73, p. 7–22. MCCAVE, I.N., 1969, Correlation of marine and nonmarine strata with
LEEDER, M.R., 1999, Sedimentology and Sedimentary Basins: Oxford, example from Devonian of New York State: American Association of
U.K., Blackwell, 592 p. Petroleum Geologists, Bulletin, v. 53, p. 155–162.
LEEDER, M.R., AND ALEXANDER, J.A., 1987, The origin and tectonic signifi- MCGOWEN, J.H., AND GARNER, L.E., 1970, Physiographic features and
cance of asymmetrical meander belts: Sedimentology, v. 34, p. 217– stratification types of coarse grained point bars: Modern and ancient
226. examples: Sedimentology, v. 14, p. 77–111.
LEEDER, M.R., MACK, G.H., PEAKALL, J., AND SALYARDS, S.L., 1996, First MCLELLAND, S.J., ASHWORTH, P.J., BEST, J.L., RODEN, J., AND KLAASEN, G.J.,
quantitative test of alluvial stratigraphy models: southern Rio Grande 1999, Flow structure and transport of sand-grade suspended sedi-
rift, New Mexico: Geology, v. 24, p. 87–90. ment around an evolving braid bar, Jamuna River, Bangladesh, in
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 167

Smith, N.D., and Rogers, J., eds., Fluvial Sedimentology VI: Interna- NORTH, C.P., AND TAYLOR, K.S., 1996, Ephemeral-fluvial deposits: inte-
tional Association of Sedimentologists, Special Publication 28, p. 43– grated outcrop and simulation studies reveal complexity: American
57. Association of Petroleum Geologists, Bulletin, v. 80, p. 811–830.
MCMECHAN, G.A., GAYNOR, G.C., AND SZERBIAK, R.B., 1997, Use of ground- OLSEN, H., 1990, Astronomical forcing of meandering river behaviour:
penetrating radar for 3-D sedimentological characterization of clastic Milankovitch cycles in Devonian of East Greenland: Palaeogeography,
reservoir analogs: Geophysics, v. 62, p. 786–796. Palaeoclimatology, Palaeoecology, v. 79, p. 99–115.
MIALL, A.D., 1986, Eustatic sea level changes interpreted from seismic OLSEN, H., 1994, Orbital forcing on continental depositional systems—
stratigraphy: a critique of the methodology with particular reference lacustrine and fluvial cyclicity in the Devonian of East Greenland, in
to the North Sea Jurassic record: American Association of Petroleum de Boer, P.L., and Smith, D.G., eds., Orbital Forcing and Cyclic
Geologists, Bulletin, v. 70, p. 131–137. Sequences: International Association of Sedimentologists, Special
MIALL, A.D., 1991, Stratigraphic sequences and their chronostratigraphic Publication 19, p. 429–438.
correlation: Journal of Sedimentary Petrology, v. 61, p. 497–505. OUCHI, S., 1985, Response of alluvial rivers to slow active tectonic move-
MIALL, A.D., 1992, Alluvial deposits, in Walker, R.G., and James, N.P., ment: Geological Society of America, Bulletin, v. 96, p. 504–515.
eds., Facies Models: Response to Sea Level Change: St. Johns, New- PAOLA, C., 2000, Quantitative models of sedimentary basin filling: Sedi-
foundland, Geological Association of Canada, p. 119–142. mentology, v. 47 (Suppl.1), p. 121–178.
MIALL, A.D., 1996, The Geology of Fluvial Deposits: New York, Springer- PAOLA, C., AND BORGMAN, L., 1991, Reconstructing random topography
Verlag, 582 p. from preserved stratification: Sedimentology, v. 38, p. 553–565.
MIDDLETON, G.V., AND SOUTHARD, J.B., 1984, Mechanics of Sediment Move- PAOLA, C., HELLER, P.L., AND ANGEVINE, C.L., 1992, The large-scale dynam-
ment: SEPM, Short Course Notes No. 3, 401 p. ics of grain-size variation in alluvial basins. 1—Theory: Basin Re-
MIKE, K., 1975, Utilization of the analysis of ancient river beds for the search, v. 4, p. 73–90.
detection of Holocene crustal movements: Tectonophysics, v. 29, p. PAOLA, C., WIELE, S.M., AND REINHART, M.A., 1989, Upper-regime parallel
359–368. lamination as the result of turbulent sediment transport and low-
MOHRIG, D., HELLER, P.L., PAOLA, C., AND LYONS, W.J., 2000, Interpreting amplitude bedforms: Sedimentology, v. 36, p. 47–60.
avulsion process from ancient alluvial sequences: Guadalope– PAOLA, C., PARKER, G., MOHRIG, D.C., AND WHIPPLE, K.X., 1999, The influence
Matarranya system (northern Spain) and Wasatch Formation (west- of transport fluctuations on spatially averaged topography on a
ern Colorado): Geological Society of America, Bulletin, v. 112, p. sandy, braided fluvial plain, in Harbaugh, J.W., Watney, W.L., Rankey,
1787–1803. E.C., Slingerland, R., Goldstein, R.H., and Franseen, E.K., eds., Nu-
MORETON, D.J., ASHWORTH, P.J., AND BEST, J.L., 2002, The physical scale merical Experiments in Stratigraphy: Recent Advances in Strati-
modelling of braided alluvial architecture and estimation of subsur- graphic and Sedimentologic Computer Simulations: SEPM, Special
face permeability: Basin Research, v. 14, p. 265–285. Publication 62, p. 211–218.
MOROZOVA, G.S., AND SMITH, N.D., 1999, Holocene avulsion history of the PAOLA, C., MULLIN, J., ELLIS, C., MOHRIG, D.C., SWENSON, J.B., PARKER, G.,
lower Saskatchewan fluvial system, Cumberland Marshes, HICKSON, T., HELLER, P.L., PRATSON, L., SYVITSKI, J., SHEETS, B., AND
Saskatchewan–Manitoba, Canada, in Smith, N.D., and Rogers, J., eds., STRONG, N., 2001, Experimental stratigraphy: GSA Today, v. 11 , no. 7,
Fluvial Sedimentology VI: International Association of Sedimentolo- p. 4–9.
gists, Special Publication 28, p. 231–249. PATRA, K.C., AND KAR, S.K., 2000, Flow interaction of meandering river
MOROZOVA, G.S., AND SMITH, N.D., 2000, Holocene avulsion styles and with floodplains: American Society of Civil Engineers, Journal of
sedimentation patterns of the Saskatchewan River, Cumberland Hydraulic Engineering, v. 126, p. 593–604.
Marshes, Canada: Sedimentary Geology, v. 130, p. 81–105. PEAKALL, J., 1998, Axial river evolution in response to half-graben faulting:
MORRIS, P.E., AND WILLIAMS, D.J., 1997, Exponential longitudinal profiles of Carson River, Nevada: Journal of Sedimentary Research, v. 68, p. 788–
streams: Earth Surface Processes and Landforms, v. 22, p. 143–163. 799.
MORRIS, P.E., AND WILLIAMS, D.J., 1999a, A worldwide correlation for PEAKALL, J., LEEDER, M., BEST, J., AND ASHWORTH, P., 2000, River response to
exponential bed particle size variations in subaerial aqueous flows: lateral ground tilting: a synthesis and some implications for the
Earth Surface Processes and Landforms, v. 24, p. 835–847. modelling of alluvial architecture in extensional basins: Basin Re-
MORRIS, P.E., AND WILLIAMS, D.J., 1999b, Worldwide correlations for sub- search, v. 12, p. 413–424.
aerial aqueous flows with exponential longitudinal profiles: Earth PEMBERTON, S.G., MACEACHERN, J.A., AND FREY, R.W., 1992, Trace fossil
Surface Processes and Landforms, v. 24, p. 867–879. facies models: environmental and allostratigraphic significance, in
MYERS, W.R.C., LYNESS, J.F., AND CASSELLS, J., 2001, Influence of boundary Walker, R.G., and James, N.P., Facies Models: Geological Association
roughness on velocity and discharge in compound river channels: of Canada, p. 47–72.
Journal of Hydraulic Research, v. 39, p. 311–320. PEREZ-ARLUCEA, M., AND SMITH, N.D., 1999, Depositional patterns follow-
NAISH, C., AND SELLIN, R.H.J., 1996, Flow structure in a large-scale model of ing the 1870s avulsion of the Saskatchewan River (Cumberland
a doubly meandering compound channel, in Ashworth, P.J., Bennett, Marshes, Saskatchewan): Journal of Sedimentary Research, v. 69, p.
S.J., Best, J.L., and McLelland, S.J., eds., Coherent Flow Structures in 62–73.
Open Channels: Chichester, U.K., Wiley, p. 631–654. PLATT, N.H., AND KELLER, B., 1992, Distal alluvial deposits in a foreland
NANSON, G.C., 1980, Point bar and floodplain formation of the meander- basin setting—the Lower Freshwater Molasse (Lower Miocene),
ing Beatton River, northeastern British Columbia, Canada: Sedimen- Switzerland: sedimentology, architecture and palaeosols: Sedimen-
tology, v. 27, p. 3–29. tology, v. 39, p. 545–565.
NICHOLAS, A.P., AND MCLELLAND, S.J., 1999, Hydrodynamics of a flood- QUINLAN, G.M., AND BEAUMONT, C., 1984, Appalachian thrusting, litho-
plain recirculation zone investigated by field monitoring and numeri- spheric flexure, and the Paleozoic stratigraphy of the eastern interior
cal simulation, in Marriott, S.B., and Alexander, J., eds., Floodplains: of North America: Canadian Journal of Earth Sciences, v. 21, p. 973–
Interdisciplinary Approaches: Geological Society of London, Special 996.
Publication 163, p. 15–26. READ, W.A., 1994, High-frequency, glacial-eustatic sequences in early
NORTH, C.P., 1996, The prediction and modelling of subsurface fluvial Namurian coal-bearing fluviodeltaic deposits, central Scotland, in de
stratigraphy, in Carling, P.A., and Dawson, M.R., eds., Advances in Boer, P.L., and Smith, D.G., eds., Orbital Forcing and Cyclic Se-
Fluvial Dynamics and Stratigraphy: Chichester, U.K., Wiley, p. 395– quences: International Association of Sedimentologists, Special Pub-
508. lication 19, p. 413–428.
168 JOHN S. BRIDGE

RETALLACK, G.J., 1997, A Colour Guide to Paleosols: Chichester, U.K., southern Utah, USA, in Flint, S., and Bryant, I.D., eds., The Geological
Wiley, 175 p. Modeling of Hydrocarbon Reservoirs and Outcrop Analogues: Inter-
RETALLACK, G.J., 2001, Soils of the past; An Introduction to Paleopedology, national Association of Sedimentologists, Special Publication 15, p.
Second Edition: Oxford, U.K., Blackwell, 404 p. 21–56.
RICE, S.P., AND CHURCH, M., 2001, Longitudinal profiles in simple alluvial SHANLEY, K.W., AND MCCABE, P.J., 1994, Perspectives on the sequence
systems: Water Resources Research, v. 37, p. 417–426. stratigraphy of continental strata: American Association of Petro-
RICHARDSON, W.R.R., AND THORNE, C.R., 1998, Secondary currents and leum Geologists, Bulletin, v. 78, p. 544–568.
channel changes around a braid bar in the Brahmaputra River, SHEAR, W.A., AND SELDEN, P.A., 2001, Rustling in the undergrowth: animals
Bangladesh: American Society of Civil Engineers, Journal of Hydrau- in early terrestrial ecosystems, in Gensel, P.G., and Edwards, D., eds.,
lic Engineering, v. 124, p. 325–328. Plants Invade the Land: New York, Columbia University Press, p. 29–
RICHARDSON, W.R.R., THORNE, C.R., AND MAHMOOD, S., 1996, Secondary 51.
flow and channel changes around a bar in the Brahmaputra River, SHEETS, B.A., HICKSON, T.A., AND PAOLA, C., 2002, Assembling the strati-
Bangladesh, in Ashworth, P.J., Bennett, S.J., Best, J.L., and McLelland, graphic record: depositional patterns and time-scales in an experi-
S.J., eds., Coherent Flow Structures in Open Channels: Chichester, mental alluvial basin: Basin Research, v. 14, p. 287–301.
U.K., Wiley, p. 519–543. SINHA, S.K., AND PARKER, G., 1996, Causes of concavity in longitudinal
ROBINSON, R.L., AND SLINGERLAND, R.L., 1998a, Origin of fluvial grain-size profiles of rivers: Water Resources Research, v. 32, p. 1417–1428
trends in a foreland basin: the Pocono Formation of the central SKELLY, R.L., BRISTOW, C.S., AND ETHRIDGE, F.G., 2003, Architecture of
Appalachian basin: Journal of Sedimentary Research, v. 68, p. 473– channel-belt deposits in an aggrading shallow sandbed braided river:
486. the lower Niobrara River, northeast Nebraska: Sedimentary Geology,
ROBINSON, R.L., AND SLINGERLAND, R.L., 1998b, Grain-size trends and basin v. 158, p. 249–270.
subsidence in the Campanian Castlegate Sandstone and equivalent SLINGERLAND, R.L., AND SMITH, N.D., 1998, Necessary conditions for a
conglomerates of central Utah: Basin Research, v. 10, p. 109–127. meandering-river avulsion: Geology, v. 26, p. 435–438.
ROBINSON, R.A.J., SLINGERLAND, R.L., AND WALSH, J.M., 2001, Predicting SLINGERLAND, R.L., AND SMITH, N.D., 2004, River avulsions and their depos-
fluvial–deltaic aggradation in Lake Roxburgh, New Zealand: test of its: Annual Review of Earth and Planetery Sciences, v. 32, p. 255–283.
a water and sediment routing model, in Merriam, D.F., and Davis, SLINGERLAND, R., HARBAUGH, J.W., AND FURLONG, K.P., 1994, Simulating
J.C., eds., Geologic Modeling and Simulation; Sedimentary Systems: Clastic Sedimentary Basins: Englewood Cliffs, New Jersey, Prentice
New York, Kluwer Academic/Plenum Publishers, p. 119–132. Hall, 220 p.
SADLER, P.M., 1981, Sediment accumulation rates and the completeness of SMITH, D.G., 1983, Anastomosed fluvial deposits: modern examples from
stratigraphic sections: Journal of Geology, v. 89, p. 569–584. Western Canada, in Collinson, J.D., and Lewin, J., eds., Modern and
SALTER, T., 1993, Fluvial scour and incision: models for their influence on Ancient Fluvial Systems: International Association of Sedimentolo-
the development of realistic reservoir geometries, in North, C.P., and gists, Special Publication 6, p. 155–168.
Prosser, D.J., eds., Characterization of Fluvial and Aeolian Reservoirs: SMITH, D.G., AND SMITH, N.D., 1980, Sedimentation in anastomosed river
Geological Society of London, Special Publication 73, p. 33–51. systems: examples from alluvial valleys near Bauff, Alberta: Journal
SANDBERG, P.A., 1983, An oscillating trend in Phanerozoic nonskeletal of Sedimentary Petrology, v. 50, p. 157–164.
carbonate mineralogy: Nature, v. 305, p. 19–22. SMITH, N.D., 1971, Transverse bars and braiding in the Lower Platte River,
SAUCIER, R.T., 1994, Geomorphology and Quaternary Geologic History of Nebraska: Geological Society of America, Bulletin, v. 82, p. 3407–3420.
the Lower Mississippi Valley: U.S. Army Corps of Engineers, Missis- SMITH, N.D., 1972, Some sedimentological aspects of planar cross-stratifi-
sippi River Commission, Vicksburg, 364 p. cation in a sandy braided river: Journal of Sedimentary Petrology, v.
SCHUMM, S.A., 1968, Speculations concerning paleohydrologic controls of 42, p. 624–634.
terrestrial sedimentation: Geological Society of America, Bulletin, v. SMITH, N.D., 1974, Sedimentology and bar formation in the upper Kicking
79, p. 1573–1588. Horse River, a braided outwash stream: Journal of Geology, v. 81, p.
SCHUMM, S.A., 1977, The Fluvial System: New York, Wiley, 338 p. 205–223.
SCHUMM, S.A., 1986, Alluvial river response to active tectonics, in Active SMITH, N.D., CROSS, T.A., DUFFICY, J.P., AND CLOUGH, S.R., 1989, Anatomy of
Tectonics: National Academy Press, Washington, D.C., p. 80–94. an avulsion: Sedimentology, v. 36, p. 1–23.
SCHUMM, S.A., 1993, River response to baselevel change: implications for SMITH, N.D., AND PEREZ-ARLUCEA, M., 1994, Fine-grained splay deposition
sequence stratigraphy: Journal of Geology, v. 101, p. 279–294. in the avulsion belt of the lower Saskatchewan River, Canada: Journal
SCHUMM, S.A., DUMONT, J.F., AND HOLBROOK, J.M., 2000, Active Tectonics of Sedimentary Research, v. B64, p. 159–168.
and Alluvial Rivers: Cambridge, U.K., Cambridge University Press, SMITH, N.D., AND ROGERS, J., eds., 1999, Fluvial Sedimentology VI: Interna-
276 p. tional Association of Sedimentologists, Special Publication 28, 478 p.
SCHUMM, S.A., AND KHAN, H.R., 1972, Experimental study of channel SMITH, N.D., SLINGERLAND, R.L., PÉREZ-ARLUCEA, M., AND MOROZOVA, G.S.,
patterns: Geological Society of America, Bulletin, v. 83, p. 1755–1770. 1998, The 1870s avulsion of the Saskatchewan River: Canadian Jour-
SCHUMM, S.A., MOSLEY, M.P., AND WEAVER, W.E., 1987, Experimental Flu- nal of Earth Sciences, v. 35, p. 453–466.
vial Geomorphology: New York Wiley, 413 p. SMITH, R.M.H., MASON, T.R., AND WARD, J.D., 1993, Flash-flood sediments
SEIFERT, D., AND JENSEN, J.L., 1999, Using sequential indicator simulation as and ichnofacies of the Late Pleistocene Homeb Silts, Kuiseb River,
a tool in reservoir description: issues and uncertainties: Mathematical Namibia: Sedimentary Geology, v. 85, p. 579–599.
Geology, v. 31, p. 527–550. SRIVASTAVA, R.M., 1994, An overview of stochastic methods for reservoir
SEIFERT, D., AND JENSEN, J.L., 2000, Object and pixel-based reservoir model- characterization, in Yarus, J.M., and Chambers, R.L., eds., Stochastic
ing of a braided fluvial reservoir: Mathematical Geology, v. 32, p. 581– Modeling and Geostatistics: American Association of Petroleum Ge-
603. ologists, Computer Applications in Geology 3, p. 3–16.
SELLIN, R.H.J., AND WILLETTS, B.B., 1996, Three-dimensional structures, STEEL, R.J., MÆHLE, S., NILSEN, H., RØE, S.L., AND SPINNANGER, A., 1977,
memory and energy dissipation in meandering compound channel Coarsening-upward cycles in the alluvium of the Hornelen basin
flow, in Anderson, M.G., Walling, D.E., and Bates, P.D., eds., Flood- (Devonian, Norway): sedimentary response to tectonic events: Geo-
plain Processes: Chichester, U.K., Wiley, p. 255–298. logical Society of America, Bulletin, v. 88, p. 1124–1134.
SHANLEY, K.W., AND MCCABE, P.J., 1993, Alluvial architecture in a sequence STOJIC, M., CHANDLER, J.H., ASHMORE, P., AND LUCE, J., 1998, The assess-
stratigraphic framework: a case history from the Upper Cretaceous of ment of sediment transport rates by automated digital photogram-
FLUVIAL FACIES MODELS: RECENT DEVELOPMENTS 169

metry: Photogrammmetric Engineering and Remote Sensing, v. 64, Yarus J.M., and Chambers, R.L., eds., Stochastic Modeling and Geo-
p. 387–395. statistics: American Association of Petroleum Geologists, Computer
STORMS, J.E.A., VAN DAM, R.L., AND LECLAIR, S.F., 1999, Preservation of Applications in Geology 3, p. 77–89.
cross-sets due to migration of current ripples over aggrading and VAIL, P.R., MITCHUM, R.M., JR., TODD, R.G., WIDMIER, J.M., THOMPSON, S., III,
non-aggrading beds: comparison of experimental data with theory: SANGREE, J.B., BUBB, J.N., AND HATLELID, W.G., 1977, Seismic stratigra-
Sedimentology, v. 46, p. 189–200. phy and global changes in sea-level, in Payton, C.E., ed., Seismic
STOUTHAMER, E., 2001, Sedimentary products of avulsions in the Holocene Stratigraphy—Applications to Hydrocarbon Exploration: American
Rhine–Meuse Delta, The Netherlands: Sedimentary Geology, v. 145, Association of Petroleum Geologists, Memoir 26, p. 49–212.
p. 73–92. VALENTINE, E.M., BENSON, I.A., NALLURI, C., AND BATHURST, J.C., 2001,
STOUTHAMER, E., AND BERENDSEN, H.J.A., 2000, Factors controlling the Ho- Regime theory and the stability of straight channels with bankfull and
locene avulsion history of the Rhine–Meuse delta (The Netherlands): overbank flow: Journal of Hydraulic Research, v. 39, p. 259–268.
Journal of Sedimentary Research, v. 70, p. 1051–1064. VAN DEN BERG, J.H., AND VAN GELDER, A., 1993, A new bedform stability
STOUTHAMER, E., AND BERENDSEN, H.J.A., 2001, Avulsion frequency, avul- diagram, with emphasis on the transition of ripples to plane bed in
sion duration, and interavulsion period of Holocene channel belts in flows over fine sand and silt, in Marzo, M., and Puidefàbregas C., eds.,
the Rhine–Meuse delta, The Netherlands: Journal of Sedimentary Alluvial Sedimentation: International Association of Sedimentolo-
Research, v. 71, p. 589–598. gists, Special Publication 17, p. 11–21.
SUN, T., MEAKIN, P., AND JOSSANG, T., 2001, Meander migration and the VAN HEIJST, M.W.I.M., AND POSTMA, G., 2001, Fluvial response to sea-level
lateral tilting of floodplains: Water Resources Research, v. 37, p. 1485– changes: a quantitative analogue, experimental approach: Basin Re-
1502. search, v. 13, p. 269–292.
SUN, T., PAOLA, C., PARKER, G., AND MEAKIN, P., 2002, Fluvial fan deltas: VAN NIEKERK, A., VOGEL, A.K., SLINGERLAND, R., AND BRIDGE, J., 1992, Routing
Linking channel processes with large-scale morphodynamics: Water heterogeneous size–density sediments over a moveable bed: model
Resources Research, v. 38, p. 1151. development, American Society of Civil Engineers, Journal of Hy-
SZERBIAK, R.B., MCMECHAN, G.A., CORBEANU, R., FORSTER, C., AND SNELGROVE, draulic Engineering, v. 118, p. 246–262.
S.H., 2001, 3-D characterization of a clastic reservoir analog: From 3- VAN TASSELL, J., 1994, Cyclic deposition of the Devonian Catskill Delta of
D GPR data to a 3-D fluid permeability model: Geophysics, v. 66, p. the Appalachian, U.S.A, in de Boer, P.L., and Smith, D.G., eds., Orbital
1026–1037. Forcing and Cyclic Sequences: International Association of Sedimen-
TEN BRINKE, W.B.M., WILBERS, A.W.E., AND WESSELING, C., 1999, Dune tologists, Special Publication 19, p. 395–411.
growth, decay and migration rates during a large magnitude flood at VEIZER, J., GODDERIS, Y., AND FRANCOIS, L.M., 2000, Evidence for decoupling
a sand and mixed sand–gravel bed in the Dutch Rhine river system, of atmospheric CO2 and global climate during the Phanerozoic eon:
in Smith, N.D., and Rogers, J., eds., Fluvial Sedimentology VI: Inter- Nature, v. 408, p. 698–701.
national Association of Sedimentologists, Special Publication 28, p. VELIKANOV, Z.M., AND YARNYKH, N.A., 1970, Field investigations of the
15–32. hydraulics of a floodplain during a high flood: Soviet Hydrology:
THORNE, C.R., RUSSELL, A.P.G., AND ALAM, M.K., 1993, Platform pattern and Selected Papers, v. 5, p. 426–440.
channel evolution of Brahmaputra River, Bangladesh, in Best, J.L., VOGEL, K., VAN NIEKERK, A., SLINGERLAND, R., AND BRIDGE, J.S., 1992, Routing
and Bristow, C.S., eds., Braided Rivers: Geological Society of London, of heterogeneous size-density sediments over a moveable bed: model
Special Publication 75, p. 257–276. verification and testing: American Society of Civil Engineers, Journal
TÖRNQVIST, T.E., 1993, Holocene alternation of meandering and anasto- of Hydraulic Engineering, v. 118, p. 263–279.
mosing fluvial systems in the Rhine–Meuse Delta (central Nether- WARBURTON, J., AND DAVIES, T., 1994, Variability of bedload transport and
lands) controlled by sea-level rise and subsoil erodibility: Journal of channel morphology in a braided river hydraulic model: Earth Sur-
Sedimentary Petrology, v. 63, p. 683–693. face Processes and Landforms, v. 19, p. 403–421.
TÖRNQVIST, T.E., 1994, Middle and late Holocene avulsion history of the WEBB, E.K., 1994, Simulating the three-dimensional distribution of sedi-
River Rhine (Rhine–Meuse Delta, Netherlands): Geology, v. 22, p. ment units in braided stream deposits: Journal of Sedimentary Re-
711–714. search, v. B64, p. 219–231.
TÖRNQVIST, T.E., AND BRIDGE, J.S., 2002, Spatial variation of overbank WEBB, E.K., 1995, Simulation of braided channel topology and topogra-
aggradation rate and its influence on avulsion frequency: Sedimen- phy: Water Resources Research, v. 31, p. 2603–2611.
tology, v. 49, p. 891–905. WEBB, E.K., and Anderson, M.P., 1996, Simulation of preferential flow in
TÖRNQVIST, T.E., WALLINGA, J., MURRAY, A.S., DE WOLF, H., CLEVERINGA, P., three-dimensional, heterogeneous conductivity fields with realistic
AND DE GANS, W., 2000, Response of the Rhine–Meuse system (west- internal architecture: Water Resources Research, v. 32, p. 533–545.
central Netherlands) to the last Quaternary glacio-eustatic cycles: WESCOTT, W.A., 1993, Geomorphic thresholds and complex response of
Global Planetetary Change, v. 27, p. 89–111. fluvial systems—some implications for sequence stratigraphy:
TUCKER, G.E., AND SLINGERLAND, R.L., 1996, Predicting sediment flux from American Association of Petroleum Geologists, Bulletin, v. 77, p.
fold and thrust belts: Basin Research, v. 8, p. 329–349. 1208–1218.
TUCKER, G.E., LANCASTER, S.T., GASPARINI, N.M., AND BRAS, R.L., 2002, The WESTAWAY, R.M., LANE, S.N., AND HICKS, D.M., 2000, The development of
Channel–Hillslope Integrated Landscape Development Model an automated correction procedure for digital photogrammetry for
(CHILD), in Harmon, R.S., and Doe, W.W., III, eds., Landscape the study of wide, shallow, gravel-bed rivers: Earth Surface Processes
Erosion and Evolution Modeling: New York, Kluwer Academic and Landforms, v. 25, p. 209–226.
Publishing, p. 349–388. WHITING, P.J., AND DIETRICH, W.E., 1993a, Experimental constraints on bar
TYE, R.S., AND COLEMAN, J.M., 1989a, Depositional processes and stratigra- migration through bends: implications for meander wavelength se-
phy of fluvially dominated lacustrine deltas: Mississippi Delta Plain: lection: Water Resources Research, v. 29, p. 1091–1102.
Journal of Sedimentary Petrology, v. 59, p. 973–996. WHITING, P.J., AND DIETRICH, W.E., 1993b, Experimental studies of bed
TYE, R.S., AND COLEMAN, J.M., 1989b, Evolution of Atchafalaya lacus- topography and flow patterns in large-amplitude meanders: 1. Ob-
trine deltas, south-central Louisiana: Sedimentary Geology, v. 65, servations: Water Resources Research, v. 29, p. 3605–3614.
p. 95–112. WILBERS, A.W.E., AND TEN BRINKE, W.B.M., 2003, The response of subaque-
TYLER, K., HENRIQUEZ, A., AND SVANES, T., 1994, Modeling heterogeneities in ous dunes to floods in sand and gravel bed reaches of the Dutch Rhine:
fluvial domains: a review of the influence on production profiles, in Sedimentology, v. 50, p. 1013–1034.
170 JOHN S. BRIDGE

WILLETTS, B.B., AND RAMESHWARAN, P., 1996, Meandering overbank flow


structures, in Ashworth, P.J., Bennett, S.J., Best, J.L., and McLelland,
S.J., eds., Coherent Flow Structures in Open Channels: Chichester,
U.K., Wiley, p. 609–629.
WILLIAMS, G.P., 1988, Paleofluvial estimates from dimensions of former
channels and meanders, in Baker, V.R., Kochel, R.C., and Patton, P.C.,
eds., Flood Geomorphology: Chichester, U.K., Wiley, p. 321–334.
WILLIS, B.J., 1989, Paleochannel reconstructions from point bar deposits: a
three-dimensional perspective: Sedimentology, v. 36, p. 757–766.
WILLIS, B.J., 1993a, Ancient river systems in the Himalayan foredeep,
Chinji village area, northern Pakistan: Sedimentary Geology, v. 88, p.
1–76.
WILLIS, B.J., 1993b, Evolution of Miocene fluvial systems in the Himalayan
foredeep through a two kilometer-thick succession in northern Paki-
stan: Sedimentary Geology, v. 88, p. 77–121.
WILLIS, B.J., 1993c, Interpretation of bedding geometry within ancient
point-bar deposits, in Marzo, M., and Puidefàbregas, C., eds., Alluvial
Sedimentation: International Association of Sedimentologists, Spe-
cial Publication 17, p. 101–114.
WILLIS, B.J., 1997, Architecture of fluvial-dominated valley-fill deposits in
the Cretaceous Fall River Formation. Sedimentology, v. 44, p. 735–
757.
WILLIS, B.J., AND BEHRENSMEYER, A.K., 1994, Architecture of Miocene
overbank deposits in Northern Pakistan: Journal of Sedimentary
Research, v. B64, p. 60–67.
WOOD, L.J., ETHRIDGE, F.G., AND SCHUMM, S.A., 1993, An experimental study
of the influence of subaqueous shelf angles on coastal plain and shelf
deposits, in Weimer, P., and Posamentier, H.W., eds., Siliciclastic
Sequence Stratigraphy: American Association of Petroleum Geolo-
gists, Memoir 58, p. 381–391.
WOODWARD, J., ASHWORTH, P.J., BEST, J.L., SAMBROOK SMITH, G.H., AND
SIMPSON, C.J., 2003, The use and application of GPR in sandy fluvial
environments: methodological considerations, in Bristow, C.S., and
Jol, H.M., eds., Ground Penetrating Radar in Sediments: Geological
Society of London, Special Publication 211, p. 127–142.
WORMLEATON, P.R., 1996, Floodplain secondary circulation as a mecha-
nism for flow and shear stress redistribution in straight compound
channels, in Ashworth, P.J., Bennett, S.J., Best, J.L., and McLelland,
S.J., eds., Coherent Flow Structures in Open Channels: Chichester,
U.K., Wiley, p. 581–608
WRIGHT, V.P., 1999, Assessing flood duration gradients and fine-scale
environmental change on ancient floodplains, in Marriott, S.B., and
Alexander, J., eds., Floodplains: Interdisciplinary Approaches: Geo-
logical Society of London, Special Publication 163, p. 279–287.
WRIGHT, V.P., AND MARRIOTT, S.B., 1993, The sequence stratigraphy of
fluvial depositional systems: the role of floodplain sediment storage:
Sedimentary Geology, v. 86, p. 203–210.
WRIGHT, V.P., AND PLATT, N.H., 1995, Seasonal wetland carbonate se-
quences and dynamic catenas: a re-appraisal of palustrine limestones:
Sedimentary Geology, v. 99, p. 65–71.
ZALEHA, M.J., 1997a, Siwalik paleosols (Miocene, northern Pakistan):
genesis and controls on their formation: Journal of Sedimentary
Research, v. 67, p. 821–839.
ZALEHA, M.J., 1997b, Fluvial and lacustrine palaeoenvironments of the
Miocene Siwalik Group, Khaur area, northern Pakistan: Sedimentol-
ogy, v. 44, p. 349–368.
ZALEHA, M.J., 1997c, Intra- and extrabasinal controls on fluvial deposition
in the Miocene Indo-Gangetic basin, northern Pakistan: Sedimentol-
ogy, v. 44, p. 369–390.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 171

ESTUARINE AND INCISED-VALLEY FACIES MODELS

RON BOYD
Earth and Ocean Sciences, University of Newcastle, New South Wales 2308, Australia
e-mail:
ROBERT W. DALRYMPLE
Department of Geological Sciences and Geological Engineering, Queens University, Kingston Ontario K7L 3N6, Canada
AND
BRIAN A. ZAITLIN
Suncor Energy Inc., Prospect Generation Services, Calgary, Alberta T2P 2V5, Canada

ABSTRACT: Modern estuaries and incised valleys are important depositional settings that have widespread significance for human land use.
The deposits of these environments are economically important for hydrocarbon exploration and production. Estuaries and incised valleys
are a complex and possibly unique environmental grouping, inasmuch as they represent creation of depositional space by one process
(mainly fluvial erosion) and fill of that space by a range of other processes (fluvial, estuarine, and marine deposition).
Early investigations of valleys began slowly in Greek and Roman times, but increased in the nineteenth century, when they were used
to develop ideas on the age of the earth in uniformitarian debates. Gradual progress was made throughout the nineteenth and twentieth
centuries with the introduction of ideas on river grade, fluvial equilibrium profiles, and base level, followed by the development of fluvial
facies models in the 1960s. Studies on estuaries began in earnest much later than those on valleys, and major advances were not made until
the mid-twentieth century, with development of the first comprehensive facies model in the 1990s.
Research on estuaries and incised valleys was energized in the 1980s by the concept of sequence stratigraphy, and work in the field has
mushroomed since then. Indeed, the currently used facies models for estuaries and incised valleys were among the first to explicitly take
into account the external control on the creation of accommodation and to be presented in a sequence-stratigraphic framework. In line with
other sedimentary environments, the facies models for estuary and incised-valley environments have also proliferated, leading to the need
for fundamental advances in how facies models are conceived.
Estuaries, as defined geologically here, are transgressive in nature. They receive sediment from both fluvial and marine sources,
commonly occupy the seaward portion of a drowned valley, contain facies influenced by tide, wave, and fluvial processes, and are
considered to extend from the landward limit of tidal facies at their heads to the seaward limit of coastal facies at their mouths. Estuaries
can be divided, on the basis of the relative power of wave and tidal processes, into two main types, wave-dominated estuaries and tide-
dominated estuaries. Estuarine facies models exhibit generally retrogradational stacking of facies and a tripartite zonation reflecting the
interaction of marine and fluvial processes. All estuaries and incised valleys have a fluvial input by definition, but estuarine facies models
reflect the balance between wave and tidal processes.
Valleys form because the transport capacity of a river exceeds its sediment supply. An incised-valley system is defined as a fluvially
eroded, elongate topographic low that is characteristically larger than a single channel, and is marked by an abrupt seaward shift of
depositional facies across a regionally mappable sequence boundary at its base. The fill typically begins to accumulate during the next base-
level rise, and it may contain deposits of the following highstand and subsequent sea-level cycles if the accommodation is not filled during
the first sea-level cycle. Incised valleys may be formed by either a piedmont or a coastal-plain river and can exhibit a simple or compound
fill. The erosion that creates many incised valleys is thought to be linked to relative sea-level fall, although climatically produced changes
in discharge and/or sediment supply may independently cause incision, even in areas far removed from the coast. In the case of valleys
in coastal areas, fluvial deposition typically begins at the mouth of the incised-valley system when sea level is at its lowest point and expands
progressively farther up the valley as the transgression proceeds, producing depositional onlap in the valley. Based on the longitudinal
distribution of broad depositional environments, the length of an incised valley can be divided into three segments. Ideally, the fill of the
seaward portion of the incised-valley (segment 1) is characterized by backstepping (lowstand to transgressive) fluvial and estuarine
deposits, overlain by transgressive marine deposits. The middle reach of the incised valley (segment 2) consists of the drowned-valley
estuarine complex that existed at the time of maximum transgression, overlying a lowstand to transgressive succession of fluvial and
estuarine deposits similar to those present in segment 1. The innermost reach of the incised valley (segment 3) is developed headward of
the transgressive estuarine–marine limit and extends to the point where relative sea-level changes no longer controlled fluvial style (i.e.,
to the landward limit of sea-level-controlled incision). This segment contains only fluvial deposits; however, the fluvial style changes
systematically due to changes in the rate of change of base level. The effect of base-level change decreases inland until eventually climatic,
tectonic, and sediment-supply factors become the dominant controls on the fluvial system. In valleys far removed from the sea, the fill
consists entirely of terrestrial deposits, but shows changes in fluvial style that are similar to those in segment 3, even though the stacking
patterns are controlled more by local tectonics and climate.
Recent and future development of estuarine and incised-valley facies models has emphasized the use of ichnology to recognize brackish-
water deposits and the ability to subdivide compound valley fills on the basis of sediment composition. Imaging the valley and its fill has
been greatly improved with 3D and 4D seismic techniques. Seabed mapping of modern estuaries has enabled detailed distributions of facies
and morphology to be compiled, enhancing the ability to predict these features in ancient rocks. Our current set of facies models represents
the early classification stage in the development of depositional models. The appropriate way forward appears to be a transformation from
qualitative approaches to empirical and quantitative computer-based models with predictive capability, based on a thorough understand-
ing of the dominant processes operating in each environment.

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 171–235.
172 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

INTRODUCTION linked facies-model approach. The other main components of


incised-valley fills (fluvial and marine sediments) are treated in
Estuaries and incised valleys and their preserved deposits separate sections of this publication.
(E&IVs; Figure 1) are important depositional settings that have Estuaries are also complex environments in that they contain
widespread significance as the sites of human habitation, as the interrelated depositional products of wave, tide, and river
harbors, as the host of significant hydrocarbon reserves, and as processes within a relatively restricted area. This complexity
the repository of important information on lowstand to early caused the development of facies models for estuaries to lag
transgressive sedimentation in ramp and shelf settings. Petro- behind those of most other adjacent environments such as rivers
leum explorationists, in particular, have focused on E&IV de- or beaches. For example, earlier editions of the Facies Models text
posits because of the economically significant quantities of (Walker 1979, 1984a, Walker and James 1992) did not have stand-
hydrocarbons produced from reservoirs hosted by the fill of alone consideration of either estuaries or incised valleys.
incised valleys (Table 1; e.g., Zaitlin and Shultz, 1984, 1990; Van Because of the nature and complexity of E&IV facies models,
Wagoner et al., 1990; Brown, 1993; Dolson et al., 1991). For this paper begins with a section on the development of both of
example, Brown (1993) estimated that ~ 25% of worldwide off- these fields, to place the concepts in their historical framework.
structure conventional petroleum traps in clastic reservoirs are The next section details the authors’ approach to facies models in
hosted within incised-valley systems, with the single largest general and the place of E&IV models in that approach. The
petroleum reserves in the world (the Athabasca Tar Sands) remainder of the paper consists of outlining the current facies
being hosted by incised-valley deposits. Therefore, a clear un- models for E&IVs, discussing how to use those models in practi-
derstanding of the internal facies architecture, reservoir charac- cal applications, illustrated by reference to both ancient and
teristics, and production behavior of incised-valley systems is of modern examples and case studies. It concludes with a section on
critical importance to the exploration for and exploitation of recent and future developments in the field.
incised-valley reservoirs.
Estuaries and incised valleys are a complex and possibly HISTORICAL DEVELOPMENT OF IDEAS ON
unique grouping of sedimentary environments, inasmuch as ESTUARINE AND INCISED-VALLEY FACIES MODELS
their formation and development involve creation of deposi-
tional space mainly by one process (fluvial erosion) and the filling Incised Valleys
of that space by a range of other processes (fluvial, tidal, and
wave), in the presence of water of variable salinity. It is the close The following discussion represents a short historical sum-
association of incised valleys with estuarine fill that has resulted mary of facies models for E&IVs. For more detail, the reader is
in these two environments being considered together here in a referred to excellent reviews such as Dalrymple et al. (1994b),

TABLE 1.—Super-giant petroleum reservoirs hosted within incised-valley (IV) deposits (i.e., reservoirs with reserves
> 50 MMBOE estimated ultimate recovery). Summarized from Dolson et al. (1991) and Pulham (1994).

Field/Trend Basin EUR (Mmboe) Age Environment


Athabasca Oil Sands Western Canada 665000 Cretaceous Fluvial–Estuarine-IV
Canada, Alberta Sedimentary Basin
Messla-Faregh Sirte 1500 Cretaceous Fluvial–Estuarine-IV
Libya
Burbank Mcalester 500 Pennsylvanian Fluvial-IV
Oklahoma
Cutbank WCSB 199 Cretaceous Fluvial-IV
Montana
Hilight Powder River 108 Cretaceous Fluvial–Estuarine-IV
Wyoming
Churches Buttes Green River 77 Cretaceous Fluvial-IV
Wyoming
South Glenrock Powder River 75 Cretaceous Fluvial–Estuarine-IV
Wyoming
Clinton Anadarko 67 Pennsylvanian Fluvial-IV
Oklahoma
Adena Denver 60 Cretaceous Estuarine-IV
Colorado
Clareton Powder River 60 Cretaceous Fluvial-IV
Wyoming
Stockholm–Arapahoe Las Animas Arch 50 Pennsylvanian Fluvial-IV
Kansas
Cusiana Llanos > 100 Eocene Fluvial–Estuarine-IV
Mirador Fm.
Colombia
ESTUARINE AND INCISED-VALLEY FACIES MODELS 173

FIG. 1.—Landsat image of the Pamlico–Albermarle Sound area of North Carolina showing a range of incised valleys, estuaries, and
lagoons. The valleys of the Pamlico, Neuse, and Roanoke rivers were incised during sea-level lowstand and have since been
flooded by relative sea-level rise. This has transformed them into wave-dominated estuaries with extensive estuary-mouth-
barrier and tidal-inlet systems. The regions between the valleys are also flooded and flanked seaward by barriers, tidal inlets, and
tidal deltas, but are better described by the term lagoons.
174 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

Miall (1996), and Blum and Törnqvist (2000), from which parts of The next major development in our understanding of rivers
the following are derived. and valleys came in the mid-twentieth century with the ideas of
The development of ideas on incised valleys is closely linked Lane (1935, 1955) and Mackin (1948), who took a more hydrody-
to the investigation of fluvial processes, and these have a long namic approach and discussed the effects of equilibrium and the
history of study back to Greek and Roman times. Plato and graded stream in terms of discharge, load, slope, and base-level
Homer both were aware of fluvial sedimentation processes. parameters. Fisk’s (1944) landmark work on the Mississippi River
Herodotus around 450 B.C. realized the connection between the developed many of these concepts into a detailed approach to a
Nile River, its valley, and the deposits at its seaward end, to single drainage system that considered its response to both
which he applied the term “delta”. The term “valley” itself internal sediment parameters and outside forcing by sea-level
derives from Latin and Old French origins meaning “a long changes. Quantitative fluvial geomorphology themes were con-
depression or hollow lying between hills or stretches of high tinued by Leopold and Wolman (1957), Leopold et al. (1964), and
ground and usually having a river or stream flowing along its later Schumm and co-workers (e.g., Schumm 1972, Schumm and
bottom” (Oxford English Dictionary). It was not until the eigh- Khan 1972, Ethridge and Schumm 1978). At around the same
teenth century that more specialized study was devoted to river time, incised-valley deposits were being recognized as hydrocar-
valleys, and this was mostly a result of the attempt to assign an bon reservoirs for the first time. One of the earliest and best-
age to the Earth. Hutton and his successor Playfair (1802) used described examples of a subsurface depositional system meeting
the idea that river valleys were the product of long-term fluvial the criteria of an incised valley was that of Harms (1966) in his
erosion to assign a much greater age to the earth than the description of stratigraphic traps within the extensive system
opposing Neptunist concepts of recent catastrophism and ori- associated with the Cretaceous “J” Sandstone in western Ne-
gin from floods. These uniformitarian themes using fluvial braska (Fig. 2).
processes were further developed in the classic work of Lyell Concurrently through the twentieth century, concepts of flu-
(1830). vial facies models were slowly being developed, beginning in the
However, it was not until later in the nineteenth century that modern sense with the work of Melton (1936), Mackin (1937), and
the concept of river grade and the fluvial equilibrium profile were Happ et al. (1940), and also in Fisk’s Mississippi studies (1944,
developed, and their relationship to valley erosion and fluvial 1947), culminating in the first major fluvial facies models devel-
deposition was appreciated. Among the first to address these oped by Allen (1963, 1964, 1965). Further developments in fluvial
concepts were Powell (1875), Gilbert (1880), who developed the facies models were summarized in the first edition of Facies
idea of base level, and Davis (1908), who illustrated the successive Models (Walker, 1979) drawing on many studies of the 1960s and
widening of a valley with age and the influence on the valley 1970s integrated in papers such as Cant and Walker (1976, 1978),
profile of strata of varying resistance. Around the same time, Miall (1977, 1978), and Rust (1978a, 1978b).
Penck and Brückner (1909) suggested a climatic control for the However, the majority of these advances did not deal with the
origin of valley terraces in southern Germany, thus initiating a longer-term evolution of river systems. Instead, they examined
continuing debate between climatic and fluvial control on valley instantaneous fluvial geomorphology, sedimentary structures,
development and fluvial deposition (e.g., Fisk 1944, 1947; Blum bedforms, paleocurrents, and empirical relationships between
1990, 1994; Blum et al. 1994; Blum and Törnqvist, 2000). parameters, finally integrating these features into static facies

FIG. 2.—Wireline-log cross section from Harms (1966) showing one of the first and best-described examples of a subsurface
depositional system meeting the criteria of an incised valley—the Cretaceous “J” Sandstone in western Nebraska. The figure
shows the J Sandstone as “valley fill”, incising the regional Skull Creek Shale and Huntsman Shale.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 175

models for meandering and braided rivers. These studies were Further advances in the recognition of incised-valley deposits
concentrated in terrestrial settings not linked to coastlines. They and documentation of fill styles were made in Van Wagoner et al.
were also concerned mainly with the detailed nature of river (1990), another book from the Exxon school of sequence stratigra-
deposits and not with a holistic approach to landscape evolution phy. The rapid acceptance of sequence stratigraphy as the pre-
that encompassed river-valley incision and the large-scale strati- ferred method for stratigraphic analysis and hydrocarbon explora-
graphic organization of deposits within those valleys. tion placed a new significance on the recognition of incised-valley
Developmental work on fluvial facies models concentrated deposits and energized the decade of the 1990s to produce the first
more on the products of deposition than erosion and hence integrated facies models for these systems. As a result, facies
moved away from the early stratigraphic emphasis on uncon- models for E&IVs were the first to explicitly include a sequence-
formities (e.g., Blackwelder, 1909; Schuchert, 1927). A change stratigraphic approach, and their usage increased rapidly (Fig. 3).
back towards these larger-scale themes was precipitated by Advances in sequence stratigraphy and its emphasis on the
the development of seismic stratigraphy and later sequence evolution of depositional systems were instrumental in the
stratigraphy in a series of papers by Vail and co-workers development of these integrated dynamic models as compared
presented first in Payton (1977) and followed up in Posamentier to the more static or autocyclic focus of earlier facies models. A
and Vail (1988) and other related papers in SEPM Special special session at the 1992 AAPG conference in Calgary was the
Publication 42 (Wilgus et al., 1988). In the latter publications, source of many of the papers that made up SEPM Special
incised valleys were seen as an integral component of a depo- Publication 51 on incised valleys (Dalrymple et al., 1994a). This
sitional sequence, formed during periods of decreasing and publication presented the first integrated facies model for an
low accommodation (e.g., Van Wagoner et al., 1990). They incised-valley system (Zaitlin et al., 1994), together with sum-
were interpreted to form by fluvial incision at the exposed maries of the history of incised-valley research (Dalrymple et
shelf break and to extend across the continental shelf and into al., 1994b) and the origin, evolution, and morphology of fluvial
the adjacent coastal plain. In this sense the application of the valleys (Schumm and Ethridge, 1994). A further 19 papers
word incised, meaning “cut into” (Oxford English Dictionary) described a range of incised-valley deposits. More recently, a
together with the word valley, was used to mean a valley that 2003 SEPM research conference on incised valleys produced an
was eroded actively as a result of allocyclic factors (particu- updated collection of research papers in another SEPM Special
larly falling relative sea level), to distinguish it from a valley Publication (Dalrymple et al., 2006).
resulting from other means (e.g., tectonic processes such as
graben formation in a rift valley; Leeder and Gawthorpe, Estuaries
1987). Thus, the criticism leveled by Blum and Törnqvist (2000)
that all valleys are incised valleys is not valid when the term is Early work on applied and environmental aspects of estuaries
used in a broad sequence-stratigraphic sense. is plentiful because of the widespread utilization of estuaries as

FIG. 3.—A search of the Georef data base (www.agiweb.org/georef) for the term “incised valley” shows a significant increase in
usage during and after the 1980s, reflecting the widespread acceptance of the sequence-stratigraphy concept (e.g., Posamentier
and Vail, 1988). Significant papers are shown in blue boxes.
176 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

harbors and ports (e.g., the Thames, the Hudson, and the Mersey; oceanographic and biological aspects, reinforced by detailed
e.g., Preddy, 1954; Hughes, 1958) or living space (the Dutch physical oceanographic studies such as Rochford’s (1951) Aus-
lowlands; Oomkens and Terwindt, 1960). The presence of tidal tralian work. Studies of ancient estuarine sediments were rare in
influence is fundamental to the concept of an estuary, and the the early twentieth century and included Arkell (1933), Pepper et
Oxford English Dictionary defines an estuary as “the tidal mouth al. (1954), and Allen and Tarlo (1963). A useful early comparison
of a great river, where the tide meets the current of fresh water” summary of modern and ancient estuarine and tidal-flat sedi-
or more fully as a “semi-enclosed coastal body of water which has ments was provided by Klein (1967).
a free connection with the open sea and where fresh water, However, while estuarine research concentrated on circula-
derived from land drainage, is mixed with sea water. Estuaries tion studies and sediment dynamics, and models for fluvial
are often subject to tidal action...” (Allaby and Allaby, 1999). systems, coasts, and deltas matured slowly, estuarine facies
However, this is primarily an oceanographic definition derived models did not make comparable advances. Schubel and
from Pritchard (1967) and is difficult to apply to sediments and Hirschberg as late as 1978 noted that “estuarine deposits rarely
ancient rocks. It can also be ambiguous in a geological context can be delimited unequivocally from other shallow marine de-
because the active progradational distributaries of a delta such as posits in the geological record because of their limited areal
the modern Mississippi can fulfill this definition of an “estuary” extent, their ephemeral character and their lack of distinctive
despite having profound stratigraphic differences from those features”. However, great strides in understanding and recogniz-
drowned (i.e., transgressive) river mouths, which are also consid- ing estuarine sediments were made from the 1960s to the present,
ered to be estuaries. such that integrated models for estuaries were finally available by
Early geological studies of the modern Severn estuary were the 1990s (e.g., Dalrymple et al., 1992).
conducted by Sollas (1883), who noted upstream sediment trans- Some of the first major steps forward in understanding the
port and determined a vertical stratigraphic succession. Other geology of estuaries were the conferences held at Jekyll Island,
early studies of estuarine sediments were conducted in the Bay of Georgia (Lauff, 1967) and Myrtle Beach, South Carolina (Cronin
Fundy (Kindle, 1917) and the Dutch estuaries and tidal flats (e.g., 1975). In the published volumes from these two conferences, an
Oomkens and Terwindt, 1960; Terwindt, 1963; Van Straaten, oceanographic definition of an estuary was formulated (Pritchard,
1952, 1954a, 1954b, 1961). Kindle also noticed upstream sediment 1967) and later largely accepted. In addition, geomorphological
transport, while Van Straaten (1952, 1954a) developed a model elements of an estuary were defined (e.g., Russell, 1967; Steers,
for tidal-channel migration (Fig. 4) well before Allen’s (1963) 1967, Jennings and Bird, 1967), rates of sediment transport and
fluvial version. Much early work seems to have differentiated accumulation were determined (e.g., Postma, 1967; Rusnak, 1967),
between tidal flats and estuaries (e.g., Klein 1967); however, and studies that indicated the tripartite sedimentary subdivision
many of the tidal-flat studies were on sediments that were com- of an estuary were presented (e.g., Kulm and Byrne, 1967; see
ponents of larger estuaries (e.g., the Bay of Fundy and the Dutch Figure 5).
and German North Sea coasts). Summaries of estuaries were Numerous studies of the morphology and evolution of tidal
produced first in the early to mid-twentieth century (Twenhofel, inlets characterized work in the 1960s and 1970s (Hoyt and
1932; Emery and Stevenson, 1957) and commonly dealt with the Henry, 1965; Vallianos, 1975; Oertel, 1975; Hine, 1975; Hubbard,

FIG. 4.—A) Block diagram and B) enlarged cross section of tidal-flat and tidal-channel sediments in the Dutch Wadden See (from Van
Straaten, 1952, 1954, as modified by Klein, 1967.)
ESTUARINE AND INCISED-VALLEY FACIES MODELS 177

FIG. 5.—Early example of tripartite estuarine sedimentation zonation, Yaquina Bay, Oregon (original from Kulm and Byrne, 1967).

1975). Many of these studies were influenced by the estuary and ment accumulated”. Most of this earlier research tended to focus
tidal-inlet ideas of M.O. Hayes, who provided the first compre- on wave-dominated rather than tide-dominated systems and on
hensive sedimentary models for these settings in his classic 1969 coastal segments that were not necessarily associated with river
and 1975 publications. Hayes (1975) also provided the basis for mouths.
division of estuaries into microtidal, mesotidal, and macrotidal Reinson’s (1992) and Dalrymple’s (1992) reviews in the third
categories, following the tidal classification system of Davies edition of Facies Models (Walker and James 1992) began to
(1964). These advances in modern systems began to be translated synthesize much of the earlier work on estuarine facies and facies
into detailed studies of ancient successions by authors such as successions and began to focus more on the role of tides. In this
Land (1972) in the Cretaceous of the Rocky Mountains, Bosence 1992 volume an early classification was developed (Fig. 6; Reinson,
(1973) in the Eocene London Basin, and Horne and Ferm (1976) in 1992), diagnostic sedimentary structures were identified, and
the Carboniferous of the Appalachians. Beginning in 1985 and summary vertical successions were provided. In addition, some
continuing through 2004, research symposia on clastic tidal sedi- integrated local studies had begun to assemble all of the basic
ments (e.g., de Boer et al., 1988; Smith et al., 1991; Bartholdy and elements required for later facies models in modern environ-
Pedersen, 2004) have provided valuable studies of many modern ments (e.g., Allen 1991; Dalrymple et al., 1990; Nichols et al.,
and ancient tidal deposits, including documentation of the tidal 1991), and in ancient rocks (Zaitlin and Schultz, 1984, 1990;
sedimentary structures by which tidal deposits can be recog- Demarest and Kraft, 1987; Rahmani, 1988; Wood and Hopkins,
nized. More recently, databases and volumes dealing with the 1989). By 1992, Dalrymple et al. had integrated many of these
distribution of estuaries across entire continents have been devel- ideas into a conceptual facies model for estuarine systems that
oped, such as those for Australia (www.ozestuaries.org) and contained a geological definition of an estuary. This work has
South America (Perillo et al., 1999). provided the main focus for research since then.
However, although extensive research continued on estuar-
ies, no comprehensive model identifying and integrating the BACKGROUND TO FACIES MODELS AND
range of geomorphological and sedimentary elements was devel- THEIR APPLICATION TO ESTUARIES
oped. Clifton’s (1982) summary catalogued many estuarine sedi- AND INCISED VALLEYS
mentary structures and proposed a tidal-channel succession. Roy
(1984) summarized much research on Australian wave-domi- Theoretical Basis of Facies Modeling
nated estuaries in a paper that identified a geomorphological
evolution that is the basis of many later models. Nichols and The facies-model concept as formulated by Walker (1984b,
Biggs (1985) provided an extensive review of estuaries, and, 1992) provides “a general summary of a depositional system
although summarizing processes and sediment dynamics com- written in terms that make the summary usable in at least (the
prehensively, noted that “it is still difficult to hindcast with following) four different ways”: (1) As a norm for comparison, (2)
certainty under what conditions and in what manner the sedi- As a framework and guide for future observations, (3) As a
178 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 6.—Early estuarine classification from Reinson (1992).

predictor in new geologic situations, and (4) As an integrated In our search for facies and geomorphological simplicity we
basis for interpretation for the system it represents. In practice, may have neglected the fundamental basis for our development
this often translated into an “idealistic” vertical succession of of models, which lies in the characteristic processes that control
facies and/or a 3D block diagram of facies relationships that sedimentation in any one depositional setting. Hence we should
supposedly portrays the “essence” of the environment. Early not expect a single model for the deltas listed above, but we
facies models had only one or a limited number of vertical should expect that all of them follow similar physical laws such
successions, and 3D block diagrams showed little internal infor- as the dispersion of suspended sediment, the response of bed
mation beyond the top and side panel(s) of the diagram. material to wave motion and the action of biological agents in the
An important extension of this approach is the display of a presence of a salinity gradient. Thus, the key to understanding
range of vertical successions in different parts of the model (e.g., depositional environments is to identify the processes that oper-
for deltas; Coleman and Prior, 1980; Galloway and Hobday, 1996) ate in each one and to determine their sedimentary response or
or a spectrum of vertical successions that illustrate the variability combination of responses. For example, the combination of waves,
that is possible, as illustrated by the multiple models for braided- longshore-directed currents, and offshore-directed rip currents
fluvial deposits by Rust (1978a), Rust (1978b), and Miall (1978), in the surf zone makes for a unique process environment. If we
summarized into the 16 “models” provided for fluvial systems by can identify the corresponding sedimentary responses and de-
Miall (1996). However, when this approach is extended to its posits for this combination of processes we will have generated a
logical conclusion, the number of “models” can proliferate, and model that summarizes those deposits and their formational
hence lose the ability to provide a relatively simple environmen- processes. It may not be the only model for nearshore marine
tal summary. In this sense the proliferation of models brings into settings, but it should be the only one that experiences that
question the provision of a “norm” (use #1 above, as discussed by specific process combination. We then need to examine the
Anderton, 1985). Part of the problem here is the degree to which physical, chemical, and biological processes of an environment,
each researcher utilizes the technique of “distillation” (Walker, as well as the properties of the sediment supplied to it, to
1984b), in which local variability is removed and replaced with a determine the range of possible outcomes for that environment.
simplified model that is based on a summary of the representa- Secondly we need to determine the probability of occurrence of
tive geomorphology and facies (i.e., an idealized view of what those process combinations and sediment types. Our ideal facies
should occur at a specific place on the earth’s surface). Hence we model then becomes one that covers the environmental range but
have problems in appreciating (for example) what the ideal view recognizes the most probable combination of processes and
of a delta is when we have to confront the contrasts between a sediments (this is the real distillation process of Walker, 1984b).
temperate-climate river-dominated mid-latitude delta and a fro- Many situations are possible in the real world, but only a small
zen arctic delta or a tide-dominated tropical delta. number are common. Environments with many variables that do
ESTUARINE AND INCISED-VALLEY FACIES MODELS 179

not display clustering of common processes and sediment types their stacking patterns. A facies model for any one environment
will not produce a single representative, useful facies model. On should take into account both the autocyclic products to provide
the other hand, the best facies models will result from environ- the building blocks and the allocyclic products that describe the
ments with few variables that exhibit frequent repetition of the geometric arrangement of those building blocks into the finished
same process combinations. Our approach to building an ideal end product.
facies model should then be a quantitative approach that models These principles can be explicitly applied to E&IV facies
the processes and sediments and is capable of creating the full models. In this case, the processes involved are primarily a
range of process–sediment interactions in an environment. Ex- combination of fluvial, wave, and tidal processes supplemented
amples of this approach are Syvitski and Daughney (1992) as locally by other processes such as organic production (e.g., peat
applied to deltas, or Cowell et al. (1992, 1995) as applied to or shell), wind, and density stratification. Wave and tidal pro-
transgressive continental shelves. Our observations derived from cesses provide a range of possibilities in estuarine systems,
experiments and field work provide the experience that identifies generating a spectrum between macrotidal, tide-dominated set-
the processes and geomorphological components, and the prob- tings and microtidal, wave-dominated settings. The combination
ability of encountering the individual examples throughout the of all E&IV processes produces a range of characteristic morpho-
range of possibilities. logical elements including river channels and flood plains,
The response to the process combination in each part of the bayhead deltas, estuarine central basins, barriers and tidal inlets,
environment will be a 3D sediment body of a particular shape that tidal deltas, and tidal sand flats and ridges. The allocyclic vari-
contains a number of characteristic properties. Sediment bodies ables produce fluvial incision during decreases in sediment in-
of this sort have been termed architectural elements (e.g., Miall, put, increases in water flux or lowered relative sea level, and
1985) and equated with facies successions by Walker (1992). The fluvial deposition followed by estuarine deposition during in-
frequent association of processes results in architectural elements creases in relative sea level and the landward migration of fluvial,
occurring in common relationships with other adjacent or linked estuarine, and marine lithofacies. Estuarine facies models are
elements. An example is the frequent association of river flood amongst the most complex due to the occurrence of multiple
plains with levees and channels because of the linked processes dominant processes (river, wave, and tide) and specific varied
of channel hydraulics and flooding. Because of the direct link responses to a range of relative sea-level and sediment-flux
between processes and facies models, the critical laboratory for parameters. This complexity contributed to the slow develop-
constructing models is the modern environment, where the inter- ment of facies models for E&IV systems.
play between sedimentary process and product can be observed
and recorded in a wide range of settings. Modern environments FORMATION AND FILL OF INCISED VALLEYS
are also becoming better suited to the documentation of sedimen-
tary architecture with the advent of high-resolution seismic sur- Incised valleys are containers. They are significant strati-
veys (particularly 3D surveys), ground-penetrating radar, graphic entities because they create a localized space in which
multibeam bathymetric surveys, and other remote-sensing tech- sediment can accumulate, often in areas where space may be
niques such as resistivity surveying. Ancient examples are not as uncommon otherwise (such as the coastal plains of low-accom-
useful because of the possibility of ambiguity in interpretation of modation basins). Incised valleys should be regarded as a system
the contemporaneous processes (the Shannon Sandstone is a in which there are two components, the valley and its fill. These
celebrated although extreme example; see for example, Suter and components may or may not be related in time or formational
Clifton, 1999) and the inability to observe those processes di- process. To understand the incision of a valley by fluvial pro-
rectly. Nevertheless, once processes have been documented and cesses (the only mechanism we will address here, neglecting
understood, observation of their depositional products in ancient valleys of structural or tectonic origin) we must consider the
rocks can be used to: (1) provide good information on the 3D sediment continuity equation , which can be written in its sim-
geometry of the deposits, (2) extend the range of variability and plest one-dimensional form as
scale for examples (such as ice-house versus hothouse climates
and a wide range of tectonic basin settings), (3) document paleo- dz/dt + dqs/dx = 0
geographic development and preservation potential through
time, as well as (4) provide the only information on non-unifor- where z = bed elevation, t = time, qs = width-averaged sediment
mitarian situations such as the pre-Silurian terrestrial processes transport rate, and x = distance along the channel. Blum and
prior to the advent of land plants and the widespread presence of Törnqvist (2000) show how this equation can be used to identify
microbial mats prior to the advent of metazoan grazers in the channel incision (an increase in z) as the result of the sediment
latest Precambrian (MacNaughton et al., 1997). transport capacity exceeding the sediment supply. Steeper slopes
The process characterization of an environment takes place at and coarser grain sizes increase the magnitude and rate of inci-
two scales, the local, autocyclic scale and the regional to global sion. Incision can result from a change in climate, tectonics, or sea
allocyclic scale. In the first case the controlling variables are level, with climate and tectonics becoming more important land-
things like fluid shear, salinity, density, and sediment size. The ward from the shoreline (Shanley and McCabe, 1994).
response is the production of distinctive sedimentary bodies that Much of the modern significance associated with incised
reflect the genetic process—these bodies are facies, facies succes- valleys derives from their association with sequence-stratigraphic
sions, and architectural elements, and the sedimentological fea- concepts (e.g., Posamentier and Vail, 1988; Van Wagoner et al.,
tures that they contain, such as bedding structures, bioturbation, 1988; Van Wagoner, 1990; Van Wagoner et al., 1991) and eco-
and their geometry. In the second (allocyclic) case, the controlling nomic importance (e.g., Brown, 1993; Dolson et al., 1991). In areas
variables are accommodation (the space made available for sedi- on the margin of a marine basin, incised valleys are considered to
mentation, sensu Jervey, 1988) and the amount and textural have formed primarily in response to a fall in relative sea level
character of the sediment flux as determined by tectonic, climatic, and a resulting decrease in accommodation, and are associated
and sea-level behavior. The sediment responses here are the with a regional unconformity. Such a response requires a spe-
production of distinctive bounding surfaces, and the generation, cific coastal-plain and continental-shelf geometry to satisfy the
preservation, and juxtaposition of stratigraphic units, including sediment continuity equation. In particular, for the sediment-
180 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

transporting capacity of the stream to increase during sea-level shoreline but lack marine influence farther inland. The facies
fall, the river must encounter a significant increase in gradient (a boundary between tidal–fluvial and purely fluvial deposits
knickpoint) somewhere seaward of the highstand shoreline (e.g., migrates landward as base level rises. Landward of the marine
Summerfield, 1985; Schumm, 1993). In other words, the fluvial limit of inundation during relative sea-level highstand, valley-
equilibrium profile lies below the level of the land surface fill deposits consist entirely of fluvial, lacustrine, and organic
(Summerfield, 1985). The incision initiated at this location then facies (e.g., Shanley and McCabe, 1994). If there is sufficient
propagates headward to create the valley. In areas with a rela- terrestrial sediment supplied during valley filling, the valley
tively low-gradient shelf and a distinct, exposed shelf–slope may be both cut and filled by fluvial processes. If the valley
break, a knickpoint generally coincides with the shelf edge. remains at least partially unfilled after sea-level lowstand, then
However, in cases without a distinct shelf break, or where the the downdip end experiences estuarine sedimentation during
shelf edge lies below the lowstand elevation, incision may not the subsequent transgression. Seaward of the highstand shore-
extend to the shelf edge; instead, recent studies of several such line, if the valley is still underfilled after transgression, some of
shelves have shown that incision begins at the break in slope the valley fill is marine and includes shelf sand and mud facies.
associated with an earlier lowstand shoreline (e.g., Woolfe et al., In valleys far removed from coastal areas, all of the valley fill is
1998; Posamentier, 2001; Fielding et al., 2003; Wellner and Bartek, fluvial in nature. Terrestrial and marine sediments are covered
2003) and/or with the immediately preceding highstand coast- in detail elsewhere in this volume and will not be considered
line. further here. Instead we will concentrate on identifying the
The lateral extent along the stream channel that can be character of estuarine sediments that are a common component
affected this way is highly debated (see Blum and Törnqvist, of valley fills in coastal areas and developing an appropriate
2000), but Quaternary examples suggest that incision across the facies model for them. Later we will return to see how estuarine
entire exposed continental shelf is possible if sea level falls sediments fit into an overall facies model for incised-valley
below the shelf edge (e.g., Suter and Berryhill, 1985), and that systems.
incision upstream of the highstand shoreline (e.g., Ethridge et
al., 1998) is possible for some tens to hundreds of kilometers: COASTAL CLASSIFICATION
Blum and Törnqvist (2000) suggest a range of from 40 to 400 km
for the upstream limit of coastal onlap. Examples of ancient Defining precisely what is or is not an estuary, and providing
incised valleys can reach hundreds of kilometers in length if the a useful geological classification scheme for estuaries, as a neces-
sea-level fall is of sufficient duration and magnitude. The Mis- sary basis for creating a facies model, has been a long-standing
sissippian Morrow Formation along the Sorrento–Mt. Pearl– problem in coastal studies. In order to solve this problem, it is first
Siaana and Stateline trends is such an example of a well-docu- necessary to present some basic ideas on coastal classification to
mented valley form that is mappable over hundreds of kilome- see what estuaries are and how they fit in (see Boyd et al., 1992,
ters (e.g., Krystinik and Blakeney-DeJarnett, 1994; Krystinik, Perillo, 1995, and Bird, 2000, for a more detailed treatment of this
1989; Bowen and Weimer, 1997). Another documented subsur- material).
face example of a long incised valley is provided by the Lower Firstly, we divide coasts into either transgressive or regressive
Cretaceous Basal Quartz and its time-equivalent units (e.g., categories (Figs. 7, 8). Secondly, we divide coasts into those that
Hayes et al., 1994; Zaitlin et al., 2002; Leckie et al., 2005). The are significantly influenced by rivers and those that are not. On
several valleys forming this compound valley fill can be traced regressive coasts, the interaction between river sediment input
for over 800 km south to north in the Western Canadian Sedi- and the ability of marine processes to redistribute that input
mentary Basin. Other examples of throughgoing valley systems determines if the coast will be an elongate or lobate protuberance
include the Pennsylvanian of the Illinois Basin (Howard and (i.e., deltaic) or linear (i.e., strandplain or shoreface or tidal flat;
Whitaker, 1988), the Permian of west-central Texas (Bloomer, Boyd et al., 1992). When the rate of relative sea-level rise exceeds
1977), the Lower Cretaceous Glauconitic Formation of Alberta the rate of sediment supply (area above the diagonal line in
(Sherwin, 1994), the Lower Cretaceous Viking–Muddy equiva- Figure 7), transgression with deposition (blue color in Figure 7)
lents in western U.S.A. (Harms, 1966; Weimer, 1984; Reinson et results in the generation of estuaries and lagoons on embayed
al., 1988; Martinsen et al., 1994; Porter and Sonnenberg, 1994), coasts and the landward migration of the shoreline and continen-
and the Upper Cretaceous Dunvegan Formation, Alberta (Plint, tal shelf on all linear (tidal-flat and headland) coasts. Coastal cliffs
2002; Plint and Wadsworth, 2003). More localized incision is fall into this latter category and form where the terrestrial gradi-
also possible: in cases where sea level does not fall very far, ent is relatively steep and there is net erosion. It is implicit in this
incision may occur only in the vicinity of the immediately arrangement that estuaries and lagoons form in areas of low
preceding highstand shoreline as a result of the relatively steep terrestrial gradient, and only during regional or local trangression.
slope of the highstand shoreface. Distinguishing such localized They should not form or persist through a shoreline regression,
incisions from tidal inlets may be difficult. In areas far removed and they should occupy only an ephemeral position at sea-level
from the sea, incision can be induced by increases in slope highstand until infilled (a critical point to appreciate for manage-
caused by tectonic activity or by an increase in the ratio of water ment of present-day highstand shorelines). However, estuaries
discharge to sediment discharge: determining the cause(s) of are commonly reestablished in the same location during subse-
incision in an ancient example can be very difficult. quent sea-level cycles, leading to multiple cut-and-fill events in
Incised-valley filling is also highly dependent on the rela- the sedimentary record. Confirmation of the formation of estuar-
tionship between accommodation and sediment flux, with fill- ies during transgressions and their disappearance during regres-
ing beginning when the fluvial equilibrium profile rises above sions is provided by the history of the 3 m sea-level oscillation of
the level of the valley base. Clearly, because valleys are incised the Caspian Sea over 65 years (Kroonenberg et al., 2000).
by fluvial processes, one can expect fluvial sediments to be Another way of describing the influence of the major coastal
deposited at the base of the valley, even if these deposits are only processes is to employ a ternary diagram identifying their rela-
one meander-belt or channel-bar height thick. In the case of tive power (Fig. 9). Here the three main process agents are
valleys cut into coastal plains, these fluvial deposits have a considered to be river currents, waves, and tidal currents. When
marine influence for some distance landward of the lowstand the ternary diagram is constructed such that the vertical axis for
ESTUARINE AND INCISED-VALLEY FACIES MODELS 181

FIG. 7.—Shoreline response (transgression versus regression) to change in sea level and sediment supply (modified from Boyd
et al., 1992).

FIG. 8.—Classification from Boyd et al. (1992), illustrating organization of all of the major clastic coastal depositional environments
based on shoreline translation direction (i.e., progradation or transgression) and relative power of waves, tidal currents, and river
currents. The upper coastline is transgressive, and the lower coastline is regressive. The influence of tides relative to wave power
increases from right to left.
182 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 9.—Triangular coastal classification using the three parameters of river, wave, and tidal processes, together with direction of
sediment supply.

fluvial power is combined with a factor that discriminates pro- geomorphologic–geologic approaches. For facies-models usage,
grading coasts from embayed transgressive coasts, and a second a geological definition is most useful because it can be applied to
factor that discriminates direct sediment supply from a river from ancient estuarine successions as well as modern estuaries. An
sediment that is supplied to the coast by marine processes (termed estuary in geological terms receives sediment from both fluvial
a marine sediment supply in Figures 8 and 9), then a clear and marine sources, commonly occupies the seaward portion of
definition of the major coastal sedimentary environments can be a drowned valley, contains facies influenced by tide, wave, and
achieved. Estuaries occupy the center of the ternary diagram, fluvial processes, and is considered to extend from the landward
where the coast is embayed and receives sediment from both limit of tidal facies at its head to the seaward limit of coastal facies
marine and fluvial sources. at its mouth (cf. Dalrymple et al., 1992). This definition overcomes
Estuaries can be distinguished by their mixed sediment source the limitations of the widely used oceanographic definition of
and association with a river input, whereas lagoons have no Pritchard (1967) based on salinity, because Pritchard’s definition
strong river-valley association and have only a marine sediment applies to both regressive and transgressive settings in addition
source (Figures 10, 11; Boyd et al., 1992;). In this scheme, estuaries to being difficult to use in ancient estuarine deposits. Estuaries as
and lagoons are intergradational, with lagoons representing the defined here are present at the mouths of valleys that are being
end-member situation where the river influence is negligible. By transgressed, and Dalrymple et al. (1992) restricted the use of
contrast, prograding deltas (the top triangle of Figure 9) derive “estuary” to such settings. However, we now recognize that
sediment directly and only from a fluvial source, whereas pro- transgressive embayments that do not contain a river-carved
grading linear coasts (strandplains and tidal flats as shown at the valley (e.g., the “abandoned” portion of a delta) may also contain
base of Figure 9) are supplied only by marine processes (waves environments that fulfill the criteria for an estuary provided
and/or tides), although that sediment must ultimately be derived above. Therefore, we extend the term “estuary” to such transgres-
mostly from a river source. It should be noted that virtually all sive settings.
coastal embayments have some form of fresh-water drainage into Most estuaries contain brackish water, but brackish water
them, making the recognition of the gradational boundary be- can occur in other settings (e.g., progradational deltas and even
tween estuaries and lagoons difficult. It is suggested here that the some shelves); hence, the identification of a trace-fossil assem-
term lagoon be used when there is no significant bedload sup- blage indicating reduced salinity in an ancient succession does
plied to the system by fluvial processes, as shown, for example, by not necessarily mean that the deposits are estuarine (sensu
the absence of a bayhead delta. Dalrymple et al., 1992). Salt-water intrusion up rivers is never as
extensive as tidal action, so an estuary as defined above extends
ESTUARINE FACIES MODEL farther inland than if a salinity-based definition is used (e.g.,
Buatois et al., 1997): the tidal limit on many modern rivers lies
Once we have identified the dominant coastal processes and tens of kilometers (in microtidal and steep-gradient settings) to
the relationship of relative sea level to sediment flux, we can hundreds of kilometers (in low-gradient, macrotidal settings)
develop a practical definition of an estuary. Perillo (1995) pro- landward of the coast.
vides an extensive discussion of estuarine definitions and classi- Because of the profound influence that waves and tides have
fications, identifying a range of oceanographic, biologic, and on their basic morphology, estuaries can be divided into two
ESTUARINE AND INCISED-VALLEY FACIES MODELS 183

FIG. 10.—A) Schematic representation of the definition of an estuary according to Pritchard (1967) and Dalrymple et al. (1992).
B) Schematic distribution of the physical processes operating within estuaries, and the resulting tripartite facies zonation.

main types, wave-dominated estuaries and tide-dominated estu- The tripartite estuarine zonation (Figs. 5, 10, 11) also corre-
aries, based on the relative power of waves and tidal processes sponds with the general patterns of net bedload transport. Long-
(Figs. 8 , 9). This distinction determines the range of the resulting term (averaged over several years) transport of bedload is sea-
facies model. Fluvial processes primarily control the upstream ward in the river-dominated zone, whereas coarse sediment
sediment flux during estuary evolution and do not alter the moves up estuary in the marine-dominated zone as a result of
fundamental morphology of the system. This point will be dis- waves and/or flood-tidal currents (Guilcher, 1967; Kulm and
cussed further in the section on criticisms, misuses, and refine- Byrne, 1967; Roy et al., 1980; Dalrymple and Zaitlin, 1989). Thus,
ments of the E&IV model. the central zone is an area of net bedload convergence and
We believe that the interaction between river and marine typically contains the finest-grained bedload sediment in the
processes provides the basis for a generalized estuarine facies estuary, regardless of whether the estuary is wave- or tide-
model. Fluvial energy, as given by the energy flux per unit cross- dominated. Once the process-based tripartite division of wave-
sectional area or other suitable measure, typically decreases and tide-dominated estuaries has been established, we can then
down an estuary (Fig. 10), because the hydraulic gradient de- examine each of these estuary types to see the major depositional
creases and the valley and its associated marine water bodies elements developed and the facies successions they produce.
widen as the river approaches the sea. Marine energy, by contrast,
generally decreases headward, because oceanic wave energy is Elements of a Wave-Dominated Estuary
dissipated by a wave-built barrier or tidal sand-bar complex and/
or because tidal-current speeds eventually decrease up the estu- The profile of “total energy” (i.e., the sum of energy from all
ary as a result of friction. Ideally, therefore, both wave- and tide- sources) for an ideal wave-dominated estuary shows two maxima,
dominated estuaries can be divided into three zones (Fig. 10): (1) one at the mouth caused by wave energy and one at the head
an outer zone dominated by marine processes (waves and/or produced by river currents, separated by a pronounced energy
tidal currents); (2) a relatively low-energy central zone, where minimum (Fig. 11). This distribution of total energy produces a
marine energy (generally tidal currents) and river currents are clearly defined, “tripartite” distribution of lithofacies (coarse–
approximately equal in strength in the long term (i.e., averaged fine–coarse) within most wave-dominated estuaries (e.g., Figs. 5,
over many years); and (3) an inner, river-dominated zone. (Note 11, 12; Kulm and Byrne, 1967; Roy et al., 1980; Zaitlin and Shultz,
that this estuarine zonation must be distinguished from the three-part 1984, 1990; Rahmani, 1988; Nichol, 1991; Nichols et al., 1991). As
segmentation of valley fills to be discussed below, because the two the estuary fills, the central energy minimum becomes less pro-
schemes have no relationship to each other). nounced.
184 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 11.—Distribution of A) energy types, B) morphological components in plan view, and C) sedimentary facies in longitudinal
section within an idealized wave-dominated estuary. MSL = mean sea level (from Dalrymple et al., 1992). Note that for simplicity
the complete transgressive succession that would be formed by landward migration of the estuary is not shown.

A marine sand body accumulates in the area of high wave the bay-head delta and the flood-tidal delta, and fine-grained,
energy at the mouth (Figs. 11, 12). It consists of a barrier, cut by organic-rich and normally bioturbated muds accumulate there
one or more tidal inlets that terminate in ebb and flood tidal (Biggs, 1967; Donaldson et al., 1970). The margins of wave-
deltas. A shoreface, which typically experiences net erosion, lies dominated estuaries typically contain salt marshes, and/or
seaward of the barrier. The limit of this shoreface or the distal mangroves cut by tidal channels, and sandy or muddy tidal
ebb-tidal delta is the marine limit of the estuary (sensu lato), and flats. A comparison of central-basin deposits between the Glau-
typically occurs in water depths less than 20 m. A subsurface conitic and Viking Formations was presented by Leroux et al.
example of such a barrier deposit located at the mouth of a (2001) and MacEachern (1999). Beaches may occur along the
wave-dominated estuary is provided by the Lower Cretaceous margins of large central basins with fetch sufficient for the local
Lloydmister Formation Senlac Pool (Zaitlin and Shultz, 1984, generation of waves.
1990), which is described below in the Incised Valley Segment 2
portion of this review. Sand and/or gravel are also deposited at Elements of a Tide-Dominated Estuary
the head of the estuary by the river, forming a bayhead delta.
This bayhead accumulation has a typical deltaic character with Most tide-dominated estuaries are macrotidal, but tidal domi-
subaerial delta plain and a subaqueous mouth bar, prodelta, nance can also occur at much smaller tidal ranges if wave action
and delta front. The morphology is typically river-dominated is limited and/or the tidal prism is large. Tidal-current energy
because of the low-energy nature of the central basin, but wave- exceeds wave energy at the mouths of tide-dominated estuaries,
and tide-dominant varieties can occur if the local processes and elongate sand bars are typically developed there (Figures
allow. A subsurface example of such a bayhead-delta deposit is 13, 14; Hayes, 1975; Dalrymple et al., 1990). These bars dissipate
the Lower Cretaceous Glauconitic Formation Lake Newell Pool the wave energy that does exist, causing it to decrease with
of Broger et al. (1997), described in the Incised Valley Segment distance up the estuary. On the other hand, the incoming flood
1 portion of this review. The low-energy central part of the tide is progressively compressed into a smaller cross-sectional
estuary (the “central basin”) acts as the prodelta region of both area because of the funnel-shaped geometry that characterizes
ESTUARINE AND INCISED-VALLEY FACIES MODELS 185

FIG. 12.—Examples of wave-dominated estuaries: A) Tuggerah Lake, NSW, on the southeast coast of Australia. Wyong Creek (right)
and Ourimbah Creek (center) are building prograding bay-head deltas into the muddy central basin of Tuggerah “Lake”, while
The Entrance tidal inlet (foreground) is building a marine sand body landward into the estuary. B) Port Stephens, NSW, showing
a merged landscape and seascape DEM illustrating the division of wave-dominated estuaries into an outer flood tidal delta and
barriers (right), a deeper central basin (middle), and an inner river valley and bay-head delta (Karuah River upper left). Depth
color bar at right is in meters below sea level.
186 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

these estuaries (Langbein, in Myrick and Leopold, 1963; Wright tide-dominated estuaries that are deeper and/or have smaller
et al., 1973). Flood tidal currents increase in speed landward tidal ranges, consists of parallel-laminated fine sand.
until frictional dissipation exceeds the effects of amplification The tripartite facies distribution is not as obvious in tide-
produced by convergence, causing the tidal energy to decrease, dominated estuaries because the energy minimum is not as
eventually reaching zero at the tidal limit. Fluvial energy de- pronounced within these channelized systems, and sands occur
creases seaward as in wave-dominated systems. The location in the tidal–fluvial channels that run along the length of the
where flood-tidal and fluvial energy are equal lies landward of estuary (Woodroffe et al., 1989; Dalrymple et al., 1990). Neverthe-
the tidal-energy maximum (i.e., the location where the tidal less, the energy minimum is the site of the finest channel sands.
current speeds are greatest; Fig. 13A). As in wave-dominated In the central, low-energy zone of systems in which the main
systems, this bedload convergence is the location of a minimum channel is unconfined by older material, this channel consistently
in the total-energy curve, but this minimum is not as pro- displays a regular progression of sinuosities (e.g., Ashley and
nounced as it is in most wave-dominated estuaries because the Renwick, 1983; Dalrymple and Zaitlin, 1989; Woodroffe et al.,
flow is channelized along the entire length of the estuary. 1989) that is termed “straight–meandering–straight” (Figs. 13,
Subsurface examples of such tide-dominated estuarine systems 15). The outer straight reach in these estuaries is tidally domi-
have been proposed from the Lower Cretaceous McMurray nated and the net sediment transport and barform asymmetry are
Formation (e.g., Flach and Mossop, 1985; Ranger and Pemberton, headward due to strong flood-tidal currents (e.g., Dalrymple et
1988) and in outcrop from the Proterozoic of Utah (Ehlers and al., 1990). The channel contains alternate, bank-attached bars (Fig.
Chan, 1999) and the Eocene of Spitsbergen (Plink-Björklund, 15B) and some mid-channel bars. The inner straight reach con-
2005). tains similar bar types, but here the net sediment transport and
In high-tidal-range end-member cases such as the Severn and barform asymmetry are downstream due to the long-term domi-
Cobequid Bay–Salmon River estuaries, the marine sand body nance of river flow over tidal currents.
consists of two strongly contrasting facies. The best known is the The region between the two straight reaches contains tight
elongate tidal sand-bar zone (Harris, 1988; Dalrymple and Zaitlin, meanders (Figs. 13, 15B) that commonly exhibit symmetrical
1989; Dalrymple et al., 1991), which is characterized by cross- point bars (Dalrymple and Zaitlin, 1989). A subsurface example
bedded medium to coarse sand (Fig. 14). These bars lie seaward of such a symmetrical tidal point-bar deposit, also from the Lower
of the tidal-energy maximum. The second facies, which coincides Cretaceous Glauconitic Formation, is provided by the Lathom
with the tidal-energy maximum, consists of upper-flow-regime “A” Pool described by Zaitlin et al. (1998). This meandering zone
(UFR) sand flats which display a braided channel pattern where is the lowest-energy portion of the system and is the position of
the estuary is broad and shallow (Fig. 15A), but these become net bedload convergence. Grain sizes in the channel become finer
confined to a single channel farther headward as the estuarine toward this area from both directions (Dalrymple and Zaitlin,
funnel narrows (Figs. 13, 15B; Hamilton, 1979; Lambiase, 1980; 1989). Muddy sediments accumulate primarily in tidal flats,
Dalrymple et al., 1990). This facies, which may not be present in marshes, and flood plains along the sides of the estuary. Subtle

FIG. 13.—Distribution of A) energy types and B) morphological elements in plan view within an idealized tide-dominated estuary.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 187

FIG. 14.—A) Overview of elongate sand bars developed in the outer (marine dominated) part of the Cobequid Bay–Salmon River
Estuary, Bay of Fundy, Canada. B) Close up of one elongate sand bar from Part A showing the scale of the bar (approximately
500 m across) and the superimposed dunes on the bar at several different length scales (Both photos by R. Dalrymple).
188 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 15.—A) The inner part of the upper-flow-regime sand flats of Zone 2, where marine energy is at a maximum (see Fig. 13). B) the
straight–meandering–straight transition in the mixed energy, upper part (Zone 3) of the Cobequid Bay–Salmon River Estuary,
Canada. This photo is taken from approximately the same position as Figure 15A but is looking in the opposite direction. The
straight channel with bank-attached bars is in the foreground, the meandering channel is in the middle distance above the bridge,
and the upper straight channel is in the upper center near the town of Truro. Both photos courtesy of John Suter.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 189

levees flank the channel, but crevasse-splay deposits become landward extent of marine influence at the time of maximum
progressively less abundant in a seaward direction through the transgression (Figs. 16, 17). In this lithosome, facies are stacked
tidal–fluvial reach because the intensity of river floods is damped retrogradationally such that the most landward terrestrial fa-
by tidal action. A discrete bayhead delta is not present in the cies is overlain by central estuarine facies and lastly by the most
river-dominated portion of tide-dominated estuaries because marine facies.
there is no open-water body into which the sediment can be The contact between the fluvial and overlying estuarine sedi-
dumped. Instead the tidal–fluvial channel passes directly into the ments is termed the initial flooding surface (FS; Figs. 16, 17), or,
river above the tidal limit. alternatively, the transgressive surface. As the estuary continues
to translate landward, the upper portion of the transgressive
Organization of Estuary Elements into a Facies Model succession is generally removed by shoreface and/or tidal-chan-
nel erosion (generating wave and tidal ravinement surfaces,
The allocyclic components of estuarine sedimentation are respectively), depending on whether wave or tidal processes
fixed, in that relative sea-level rise over the long term exceeds dominate. The amount of section removed varies between ex-
the sediment input from both marine and fluvial sources, result- amples, depending on the relationship between the rates of sea-
ing in transgression, a necessary condition for the formation of level rise and transgression, the rate of sediment input, the depth
estuaries (as defined geologically; Dalrymple et al., 1992). Estu- of the shoreface and tidal-channel thalweg, and the depth of the
aries are typically initiated with the beginning of the transgres- paleovalley (cf. Davis and Clifton, 1987; Demarest and Kraft,
sion and continue accumulating sediment throughout the trans- 1987). Partial transgressive successions, in which the basal fluvial
gression, up to the time of maximum flooding, when the shore- and fluvial–estuarine facies have the highest preservation poten-
line reaches its most landward position, before finally filling at tial, should occur along the transgressed portion of a paleovalley,
the beginning of the subsequent highstand. If the highstand is of seaward of the highstand shoreline (Figs. 16, 17). Fluvial deposits
short duration, sea level may fall before the estuary is com- should occupy the deepest portions of the valley, except near the
pletely filled; however, if the highstand is long and/or the rate lowstand river mouth, where tidal–fluvial sediments may occur.
of sediment input is high, then the estuary fills completely in the Along the flanks of the valley, estuarine deposits lie directly on
transition to highstand progradation. As a result, an assemblage older deposits and the sequence boundary, without intervening
of estuarine facies, termed here an estuarine lithosome, stretches fluvial sediments. In settings where estuaries occupy embayments
along a substantial portion of the valley or the length of the that are not paleoriver valleys, the estuarine deposits overlie
embayment, from near the lowstand mouth of the river to the either earlier deposits such as deltas and are separated by a

FIG. 16.—Schematic section along the axis of a wave-dominated estuary, showing the distribution of lithofacies resulting from
transgression of the estuary, followed by estuary infilling and shoreface progradation at the time of sea-level highstand. The
completeness and thickness of the preserved transgressive succession depends on the relative rates of sea-level rise and the
headward translation of the shoreface. See Figure 17 for legend (from Zaitlin et al., 1994). “Flooding surface (FS)” refers to the
initial flooding surface at the beginning of transgression.
190 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 17.—Schematic section along the axis of a tide-dominated estuary, showing the distribution of lithofacies resulting from
transgression of the estuary, followed by estuary infilling and progradation of sand bars or tidal flats. The completeness and
thickness of the preserved transgressive succession depends on the relative rates of sea-level rise and the headward translation
of the thalweg of tidal channels (from Zaitlin et al., 1994). “Flooding surface (FS)” refers to the initial flooding surface at the
beginning of transgression.

flooding surface, or older unrelated units located below an un- ter buildups, may also be present at this stratigraphic level. The
conformity. finest sediments are overlain in turn by an upward-coarsening
succession passing into either flood-tidal delta and washover
Wave-Dominated Estuarine Model sediments (Fig. 16, C1, C2, C3) along most of the length of the
estuarine lithosome, or into bayhead-delta deposits (Fig. 16, C3)
The marine sand body in these estuaries is a composite feature at locations where there is episodic bayhead-delta progradation.
that may contain several discrete facies. In transgressive succes- Tidal-channel migration during transgression generates a tidal
sions, some or all of the barrier complex is likely to be eroded ravinement surface landward and ahead of the wave ravinement
during shoreface retreat and overlain by a wave ravinement surface, providing at least two possible erosion surfaces within
surface (Fig. 16, C1). If any part of the barrier remains, it consists the wave-dominated estuarine succession.
of the deeper and/or more landward facies including erosionally The bayhead delta deposits are distinguished from true flu-
based tidal-inlet deposits and the landward-directed cross bed- vial sediments by the presence of tidal structures and/or a
ding of washovers and flood-tidal deltas that may interfinger brackish-water fauna as well as a deltaic geometry and stratigra-
with the underlying central-basin muds (e.g., Roy et al., 1980; phy. Bayhead-delta sediments are likely to be common in the
Roy, 1984; Zaitlin and Schulz, 1984, 1990; Boyd and Honig, 1992). lower part of transgressive valley-fill successions, and will occur
In vertical profile, fine-grained central-basin sediments ideally at the up-dip end of the estuarine lithosome where they will
exhibit a symmetrical grain-size trend (Fig. 16, C4). The basal exhibit an upward-coarsening succession resulting from progra-
upward-fining portion represents the passage from transgres- dation either during stillstands or during estuary filling at high-
sive, fluvial, and bayhead-delta deposits through progressively stand (Fig. 16, C4; Rahmani, 1988; Reinson et al., 1988; Allen, 1991;
more distal prodelta sediments. More commonly, the base of the Allen and Posamentier, 1993; Broger et al., 1997).
central-basin muddy facies is an abrupt flooding surface that Meandering tidal channels containing inclined heterolithic
might display some evidence of erosion (i.e., a “bay ravinement strata (Flach and Mossop, 1985; Thomas et al., 1987; Pemberton
surface”) that occurred as the low-energy central-basin shoreline and Wightman, 1992) are likely to be most abundant in the late
transgressed. The finest sediments represent the center of the stage of estuary filling, when the prograding bayhead delta
central basin and are frequently the mostly intensely bioturbated merges with the flood-tidal delta (Smith, 1987; Nichol, 1991).
(although often with an impoverished, brackish-water trace- Such channels may erode some or all of the underlying central-
fossil assemblage). Organic facies, including peat, coal, and oys- basin succession and might scour down to the basal unconfor-
ESTUARINE AND INCISED-VALLEY FACIES MODELS 191

mity. An additional stratigraphic surface, termed the bayhead- depositional facies across a regionally mappable sequence bound-
delta diastem, may be generated by erosion at the base of ary at its base. The fill typically begins to accumulate during the
laterally migrating bayhead-delta distributaries (e.g., Nichol, next base-level rise, and may contain deposits of the following
1991). Ancient wave-dominated estuarine systems such as the highstand and subsequent sea-level cycles” (Zaitlin et al., 1994).
Lower Cretaceous Lloydminster Member and the Albian Paddy
Member (Leckie and Singh, 1990; Leckie et al., 1990) will be Types of Incised Valleys
discussed in the later section dealing with incised valleys,
segment 2. There are two major physiographic types of incised valley.
Incised-valley systems that have their headwaters in a (moun-
Tide-Dominated Estuary Model tainous) hinterland, and that cross a “fall line” where there is a
significant reduction in gradient, are here considered to be pied-
During transgression, the elongate tidal sand bars that consti- mont incised-valley systems. There are many ancient examples
tute the outer part of the marine sand body in tide-dominated from the North American Western Interior Seaway that can be
estuaries are likely to be erosionally truncated or completely interpreted as piedmont incised-valley systems including the
removed (Fig. 17, C1) by the headward migration of the erosional Lower Cretaceous Cutbank, Taber, and Basal Quartz of northern
zone that coincides with the “bedload parting” that lies seaward Montana–Alberta (e.g., Hayes, 1986; Dolson and Piombino, 1994;
of the estuary mouth (Dalrymple, 1992; Dalrymple et al., 1992) Ardies et al., 2002; Lukie et al., 2002; Zaitlin et al., 2002), Glauco-
and/or the headward and lateral migration of tidal channels that nite Formation (Rosenthal, 1988; Sherwin, 1994) and Muddy
separate the sand bars. The amalgamation of these scours pro- Sandstone and its Canadian equivalents the Viking and Bow
duces a tidal ravinement surface. Erosion by the channels during Island Formations (Gustason et al., 1986; Dolson et al., 1991;
transgression also causes the cross-bedded sand bars, or the Pattison, 1991; Pattison and Walker, 1994, 1998; MacEachern and
parallel-laminated, UFR sand-flat deposits, to overlie or abut Pemberton, 1992, 1994). Incised-valley systems that are confined
erosionally against mudflat and salt-marsh sediments along the to low-gradient coastal plains and that do not cross a “fall line”
margins of the estuary and/or on more headward facies in the are termed coastal-plain incised-valley systems. Subsurface ex-
axis of the valley (Fig. 17, C2). If the transgressive succession amples of coastal-plain estuaries include parts of the Cretaceous
contains both sandy facies (i.e., cross-bedded medium to coarse Viking Formation (e.g., Pattison, 1991; MacEachern and
sand and parallel-laminated fine to very fine sand), they produce Pemberton, 1994) at Sundance, Edson, and CynPem, and the
an overall upward-coarsening trend. The contact may be either southern portions of the Paddy–Cadotte (e.g., Leckie and Singh,
erosional or gradual. 1991).
The central, mixed-energy (tidal–fluvial meanders) and in- Piedmont incised-valley systems are characterized by a longer
ner, river-dominated portions of the estuary are characterized by fluvial reach than coastal-plain systems and are commonly asso-
tidal-channel deposits that are flanked by vertically accreted, ciated spatially with underlying structural features in the hinter-
salt-water, brackish-water, and fresh-water marsh sediments. If land, e.g., the Upper Cretaceous Dunvegan System (Plint, 2002,
sufficient accommodation is generated, the point-bar sediments and the Mississippian Morrow System (Bowen and Wiemer,
of the meandering zone are overlain and underlain by the depos- 1997, 2003). As a result, these river systems may be longer lived
its of straighter channels (Fig. 17) that display opposite paleocur- than coastal-plain systems. Also, piedmont systems more com-
rent directions; if there is low accommodation, the last channel to monly contain coarse-grained, less-mature, fluvially supplied
cross the area incises into the older tidal-channel deposits. Upper- sediment, whereas coastal-plain systems are usually filled by
flow-regime parallel lamination predominates in the shallower finer-grained and more mature deposits recycled from coastal-
parts of the outer (tide-dominated) straight reach (Fig. 15A), plain sediments. Piedmont systems may have overall higher rates
while dunes may occur in the deeper channels. Ripples and/or of sediment supply because they have larger catchment areas. In
dunes are likely to be more abundant in the meandering and both piedmont and coastal-plain systems, marine-derived sedi-
inner straight reaches. The channel sediments are finest, and the ment is preserved in the estuarine portion of the valley fill (see
mixing of fluvially and tidally supplied sediment is most pro- below). Coastal-plain and piedmont incised-valley systems oc-
nounced, in the zone of tight meandering. The contacts between cur adjacent to each other in modern coastal areas (e.g., Hayes and
facies zones coincide with erosional channel bases. The channel- Sexton, 1989).
bank sediments consist of tidally bedded sands and muds that
occur either as erosionally bounded wedges of flat-lying strata Simple and Compound Incised-Valley Fills
(Dalrymple et al., 1991) or as inclined heterolithic strata (IHS); see
Flack and Mossop (1985). IHS is most prevalent in the meander- The fill of any incised-valley system can be classed as either
ing reach. A well-documented example of an ancient tide-domi- simple or compound depending on the absence or presence,
nated estuary is the Cretaceous Lower Greensand in the Leighton respectively, of multiple, internal, high-frequency sequence
Buzzard area of England (Johnson and Levell, 1995). boundaries. If the valley is filled completely during one cycle
such that the depositional surface rises above the level of the
INCISED-VALLEY FACIES MODEL original interfluves, the fill is termed a “simple fill”. An ancient
example of a simple fill has been described by Zaitlin and
To develop an appropriate facies model for an entire incised Schultz (1984, 1990; see more below). A “compound fill” records
valley, compared to an estuary, we need to address the wider multiple cycles of incision and deposition resulting from fluc-
concept of an incised-valley system (IVS). An incised-valley tuations in base level and is therefore punctuated by one or
system (e.g., Fig. 18) must incorporate elements of the erosional more sequence boundaries in addition to the main, lower-order
valley itself, the strata that it eroded into, and the entire fill sequence boundary at the base of the incised valley (e.g., the
consisting of fluvial, estuarine, and marine facies (Fig. 19). In this Mississippian Morrow Formation; Krystinik and Blakeney-
context, an incised-valley system is defined as “a fluvially eroded, DeJarnett, 1994; Krystinik, 1989; Bowen and Weimer, 1997,
elongate topographic low that is typically larger than a single 2002); and the Lower Cretaceous Basal Quartz Formation (Ardies
channel form, and is characterized by an abrupt seaward shift of et al., 2003; Zaitlin et al., 2002; Leckie et al., 2005), the Lower
192 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 18.—Example of incised valley with incised tributaries, Red Deer River south of East Coulee, Alberta, Canada. An incised-valley
system consists of the erosional form seen here, plus the sediments that will ultimately fill this container.

Cretaceous Glauconitic Formation (e.g., Wood and Hopkins, larger than a single channel (e.g., Figs. 18, 23) and commonly
1989, 1992; Broger et al., 1997), and the Viking/J/Muddy For- has an erosional relief (from the valley base to the original
mation (Gustason et al., 1986, Gustafson et al., 1988, Reinson et floodplain level) of 10 m or more. However, there is a com-
al., 1988). Due to the presence of structural control on their plete gradation from non-incised channels, through shal-
location, piedmont river systems may exist through more than lowly incised systems, to very deeply entrenched valleys
one sequence of sea-level fall and rise; thus, their incised valleys (Fig. 23). Studies of both modern and ancient valleys show
may contain a compound fill, although higher rates of sediment that the depth of incision is not constant along their length
supply may counteract this tendency (e.g., Gustason et al., 1988; (Schumm and Ethridge, 1994). Deeper-than-average incision
Dolson et al., 1991; Ardies et al., 2002; Zaitlin et al., 2002). occurs at the location where tributaries join the trunk river
Coastal-plain systems are more likely to exist through only one (scour depths at these locations may be up to five times the
regression–transgression cycle and therefore have a simple fill, depth of adjacent parts of the valley; Best and Ashworth,
unless the rate of sediment supply is too low to fill the valley 1997), at the location of flow constrictions where the river cuts
during a single cycle. across a more resistant underlying unit, and at the outsides of
bends. Ardies et al. (2002) show a well-documented ancient
Recognition of Estuarine and Incised-Valley Systems example of all three types of channel-bottom irregularity (Fig.
24). The valley width may also be quite variable; it increases
E&IV deposits are among the hardest to recognize because of with time (e.g., Schumm and Ethridge, 1994) and is wider
their low width:depth ratio, limited lateral extent and ribbon where the river cuts into less resistant lithologies (e.g., Ardies
geometry, and the complex association of fluvial, tidal, wave, and et al., 2002). However, typical dimensions are in the range of
marine facies within them (Figs. 19–21). The following is a list of several hundreds of meters to several tens of kilometers, with
criteria for recognizing E&IV systems: most valleys in the range of 1–10 km wide.

(1) The valley is a negative (i.e., erosional) paleotopographic (2) The base and walls of the incised-valley system represent a
feature, the base of which truncates underlying strata, sequence boundary (Fig. 22, red line) that correlates to an
including any regional markers (such as bentonites, coals, erosional (or hiatal) surface outside the valley (i.e., on the
flooding surfaces, or seismic markers) that may be present interfluve areas). This erosional surface may be modified by
(Fig. 22, green arrow). The valley container has a characteris- later transgression, forming an E/T (erosive–transgressive)
tic size, shape, and regional extent. The valley should be surface; Plint et al., 1992), or a combined flooding surface and
ESTUARINE AND INCISED-VALLEY FACIES MODELS 193

sequence boundary (an FS/SB surface; Van Wagoner et al., unincised channels and augment criterion 2 above. On the
1990). The sequence boundary may be mantled by a pebble regional scale, the planform geometry of tributary networks
lag and/or characterized by burrows belonging to the should mimic the river system(s) that became entrenched. As
Glossifungites ichnofacies (MacEachern et al., 1992; a result, the various river patterns identified by geomor-
MacEachern and Pemberton, 1994). On the interfluves the phologists (e.g., Howard, 1967) may be recognized in valley
exposure surface may be characterized by a particularly systems. For example, dendritic patterns predominate in
well-developed soil or rooted horizon (Leckie and Singh, areas with uniform slopes and substrate erodibility, whereas
1991; McCarthy and Plint, 1998). Such paleosols may show rectilinear patterns occur in jointed bedrock or in areas with
evidence of lower groundwater tables and more prominently a crosscutting network of subtle faults. Recent work suggests
developed soil horizons than paleosols formed that faults that are active during incision may have a strong
syndepositionally within the TST and/or the HST. influence on the location and planform pattern of valleys
(e.g., Ardies et al., 2002).
(3) Because the river erodes below the level of the interfluves
when it creates the valley, water drains downward into the (4) A fundamental aspect of incised valleys is their formation at
valley; as a result, the trunk river may be fed by smaller times of erosion and falling base level; in cases where the
incised tributary valleys that are themselves incised (e.g., area lies close to the shoreline, coastal regression accompa-
Figs. 18, 24; Posamentier, 2001; Ardies et al., 2002). These nies incision. Hence, the base of the incised-valley fill (Figs.
tributary valleys aid in distinguishing incised valleys from 19–21) exhibits an erosional juxtaposition of more proximal

FIG. 19.—Idealized longitudinal section of a simple incised-valley system showing the distribution of A) depositional environments,
B) systems tracts, and C) key stratigraphic surfaces. A wave-dominated estuary has been used in this model. Segments 1 and 3
are typically much longer than segment 2, and are compressed here for presentation purposes. Also shown are the locations of
the schematic profiles illustrated in Figure 20. Modified from Zaitlin et al. (1994).
194 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 20.—Five representative vertical sections of facies and sequence-stratigraphic surfaces in an idealized incised-valley system,
based on an estuarine system that is wave dominated. WRS = wave ravinement surface, MFS = maximum flooding surface, IFS
= initial flooding or transgressive surface, SB = sequence boundary, TRS = tidal ravinement surface, BHD = bay-head delta.
Numbers in circles identify location of sections shown in Figure 19. Modified from Zaitlin et al. (1994).

(landward) facies over more distal (seaward) deposits (i.e., valley, or of each sequence constituting a compound valley
a “basinward shift in facies” sensu Van Wagoner et al., fill. A maximum flooding surface lies above the valley fill in
1990), across a regional hiatus (unconformity)—vertical segment 1, within the estuarine deposits in segment 2, and
white arrow in Figure 22. The subsequent filling of the likely low in the fluvial deposits in segment 3. Wave and tidal
valley occurs partially or wholly during rising base level ravinement surfaces are commonly present between the se-
and is accompanied by transgression in near-coast situa- quence boundary and the maximum flooding surface in the
tions. The latter typically results in more downdip facies areas transgressed by the shoreline. Additional flooding
(marine, estuarine) being deposited on top of more updip surfaces, bay ravinement surfaces, and erosional surfaces of
facies (terrestrial). In the case of valley fills consisting solely more local extent, including bayhead and fluvial diastems,
of fluvial facies, those facies reflect the change from a low- are likely to be formed during backstepping of fluvial and
accommodation to a higher-accommodation style, for ex- estuarine subenvironments.
ample by changing the channel stacking patterns, the rela-
tive preservation of overbank deposits, or the amount of (7) Channels contained within the valley should be substantially
organic facies (Fig. 21), and/or by a change in any paleosols smaller than the valley itself (e.g., Figs. 18, 21, 23). However,
from well-drained and more mature to poorly drained and it is recognized that channels that experienced only a short
immature as accommodation increases. period of incision may be incised only slightly, with insuffi-
cient widening to form a pronounced valley. In addition, as
(5) As a result of filling in response to rising base level, deposi- discussed above, individual scours within a channel may be
tional markers within the deposits of the incised-valley fill much deeper than the average channel depth, for example at
onlap the valley base and walls but do not occur outside the tributary junctions (e.g., Best and Ashworth, 1997; Ardies et al.,
valley (smaller white horizontal arrow in Figure 22), except 2002). In these cases, the deeper scour could be mistaken for a
where they can be traced in a seaward direction into equiva- valley but is of local extent only (Fig. 24), whereas a valley
lent marine deposits. exhibits an elongate erosion surface of more regional extent.
Where the valley and channel boundaries can be observed
(6) In terms of sequence-stratigraphic surfaces (Figs. 19–21), the together, floodplain or terrace surfaces attached to channels
formation of a valley generates a sequence boundary at the within the valley can occur at lower stratigraphic elevations
base, and a transgressive surface within the fill of a simple than the adjacent valley walls (M. Boyles, personal communi-
ESTUARINE AND INCISED-VALLEY FACIES MODELS 195

FIG. 21.—Nonmarine sequence-stratigraphic model showing the change in channel stacking patterns and organic facies responding
to a cycle of accommodation change such as may be seen in segment 3 of an incised-valley fill. From Boyd and Diessel (1994).

cation, 2002) and/or interfluves outside the valley. This recog- nistic behavior (cf. Howard and Frey, 1973, 1975; Howard et
nition feature augments those listed in 4 and 5 above. al., 1975; MacEachern and Pemberton, 1994; Buatois et al.,
1997; Buatois et al., 2005; Gingras et al., 1999; Pemberton et
(8) Estuaries as defined above, following Dalrymple et al. al., 2001). The degree of bioturbation (i.e., the bioturbation
(1992), are transgressive, tidally influenced environments index; Droser and Bottjer, 1986, 1989) is commonly highly
that constitute an important and distinctive component of variable, with essentially unbioturbated beds interbedded
incised valleys in their seaward parts. Because estuaries with extensively bioturbated deposits that may contain a
tend to enhance tidal action because of flow constriction, tidal monospecific assemblage of traces. The unconformity at the
indicators and distinctively tidal deposits may be especially base of the valley can display a Glossifungites ichnofacies
abundant within the fill of incised valleys. The most distinc- (MacEachern et al., 1992), and individual forms such as
tive of these (Fig. 25) are flat-lying tidal rhythmites and tidal Gyrolithes are distinctive of the estuarine environment (see
bundles in cross beds, both of which record the neap–spring the Brackish Ichnology section below). Brackish-water mi-
tidal cyclicity. In more general terms, single and paired (i.e., crofauna and macrofauna also display distinctive diversity
double) mud drapes, which give the deposits a heterolithic and occurrence trends that are useful for the recognition of
nature, are indicative of tidal sedimentation. In addition, estuarine deposits, such as marsh foraminifera (Ammonia,
other classic features such as reactivation surfaces, bidirec- Haplophragmoides, Trachammina sp.) that occur primarily in
tional paleocurrent patterns, herringbone cross stratification, the intertidal zone in combination with Spartina sp. flora.
flaser to lenticular bedding, and the large scale of cross-beds Bivalves such as the modern Rangia cuneata that are over-
are distinctive (Dalrymple, 1992). In relatively low-accom- whelmingly found in estuarine settings and oysters such as
modation settings and in basins with small tidal ranges, Crassostrea sp. are also useful environmental indicators.
incised-valley fills may be the only place where tidal deposits However, it is important to note that many of these brack-
are preserved. In such cases, the presence of tidal deposits can ish-water features may occur in settings other than estuar-
be used to suggest the existence of an incised valley. ies and should not be used on their own to interpret the
presence of an estuary or an incised valley.
(9) The mixing of fresh and salty water is a fundamental
characteristic of estuaries. This stressed environment pro- (10) Estuaries receive sediment input from both the marine and
duces a characteristic ichnological suite and faunal compo- terrestrial ends of the system (Figs. 9–12), creating the poten-
sition (Pemberton et al., 1992, and early articles in Lauff, tial for the mixing of sediment with two different composi-
1967) that are characterized by a low ichnospecies diversity, tions. The sediment supplied directly by the river reflects the
with populations consisting of small individuals (smaller bedrock composition of the fluvial drainage basin, while the
then their open-marine counterparts) that exhibit opportu- sediment provided by the marine source reflects shelf litholo-
196 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 22.—The four main criteria for recognizing an incised-valley system illustrated using a Lower Cretaceous (Basal Quartz
equivalent), muddy incised-valley fill cutting into shoaling-upward shelf–shoreface parasequences along the Missouri River
in northern Montana, U.S.A. Photo courtesy of P. Putnam, Petrel Robertson Research.

gies and/or source regions updrift in the longshore transport


system. Because the marine-sourced material has been re-
worked from older deposits, it is often more mature than the
terrestrial sediment (e.g., Roy, 1977). If source regions change
during the fill of a compound valley, it may be possible to
distinguish individual sequences within the compound val-
ley fill by their compositional differences (e.g., Zaitlin et al.,
2002).

(11) E&IVs contain a characteristic mix of sedimentary facies.


These include terrestrial, estuarine, and marine facies and
range from fluvial, to tidal–fluvial channel, bayhead delta,
central basin, barrier, and tidal sand ridge. When found in
combination, and especially when such facies are not present
in the surrounding regional deposits, this set of facies iden-
tifies an estuarine setting, provided that they display a
transgressive stacking arrangement (Figs. 16, 17) and may
also point to the presence of an incised valley if a suitable
container is present. Note that the presence of fluvial facies
at the base of the estuarine or valley-fill succession is helpful
but not essential for identification. Transgression subse-
quent to fluvial deposition can result in reworking and
removal of fluvial facies by tidal and wave processes, espe-
cially by means of erosion at the bases of migrating tidal
inlets. In other situations, the fluvial sediments may not be
widespread and may occur only in a geographically re-
stricted zone along the valley axis. Near the seaward end of
FIG. 23.—Incised-valley formation and entrenchment. If the flood- segment 1 of the incised-valley system, all channel facies are
plain is periodically inundated by large floods, the river is not likely to be tidal–fluvial in nature and hence display tidal
incised, regardless of the relief between the low-stage water features.
level in the river and the floodplain. The situation shown in
Part B is the minimum incision required to qualify the river as (12) The central zone of incised-valley estuaries is occupied by
incised in the modern, but such situations may be difficult to a low-energy region (Figs. 11, 13) representing either the
distinguish from non-incised channels in the ancient; the finer-grained central basin of wave-dominated estuaries or
degree of development of floodplain paleosols, if preserved, the fine-grained meandering reach of tide-dominated estuar-
would be the key distinguishing factor. ies.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 197

FIG. 24.—A 3-D amplitude anomaly map of a part of the Basal Quartz Formation (Lower Cretaceous) of southern Alberta, showing
a tributary-junction scour (TJS) (cf. Ardies et al., 2001; 3-D image courtesy PanCanadian Energy (now EnCana) Corporation).

(13) In the case of valley incision during regression and relative ful tools for the identification of subtle structural warping and/
sea-level fall due to steepening of the fluvial profile as a result or fault movement (e.g., Ardies et al., 2002), because rivers seek
of seaward extension of the river, the regional marine gradient out the lowest part of the eroding landscape. A variety of tech-
is greater than the terrestrial gradient of the river valley. niques have been employed to identify and map paleovalleys,
including: (1) geological structure mapping of the erosional sur-
(14) E&IV deposits occupy fluvial drainage corridors, and their face from 2D–3D seismic (e.g., Broger et al., 1997) or from wireline
locations are often determined by underlying paleotopo- logs (e.g., Krystinik, 1989; Van Wagoner et al., 1990; Krystinik and
graphic and structural trends, with valleys occurring espe- Blakeney-DeJarnett, 1994; Bowen and Weimer, 1997); (2) third- or
cially in areas of subtle downward flexure and/or parallel to higher-order residual mapping of the erosional surface in areas
fault traces (cf. Ardies et al., 2002; Plint and Wadsworth, 2003). affected by postdepositional structuring (e.g., Zaitlin and Shultz,
By contrast, valleys tend to avoid areas of subtle upwarping. 1984, 1990); (3) detailed isopach mapping of the interpreted fill, or
of an interval between the unconformity and an overlying hori-
An early example of an interpreted subsurface incised-valley zontal marker that extends over the interfluves, to locate anoma-
system that subsequently met many of the above criteria for an lously thick sections confined to the paleotopographic lows (e.g.,
incised valley was that of Harms (1966) in his description of the Siever, 1951; Van Wagoner et al., 1990, Ardies et al., 2002). Other
Cretaceous “J” Sandstone in western Nebraska. Harms’ correla- techniques include petrographic and chemostratigraphic typing
tion (Fig. 2), based on a detailed electric-log cross section, demon- of sediment composition, gravity techniques, resistivity map-
strated the truncation of regionally mappable, coarsening-up- ping, and mapping of hydrocarbon production trends.
ward marine parasequences by blocky to fining-upward fluvial
valley-fill deposits, thus fulfilling recognition criteria (1), (2), (4), Model for a Simple Incised-Valley Fill
and (5) above. Other examples include those of the Mississippian
Morrow Formation (e.g., Krystinik, 1989; Krystinik and Blakeney, For simplicity, here we present a model for a simple incised-
1990; Krystinik and Blakeney-DeJarnett, 1994; Bowen and Weimer, valley fill, based primarily on Zaitlin et al. (1994). We will con-
1997, 2003) and parts of the Lower Cretaceous Glauconitic Forma- sider the case of a piedmont incised-valley system, which is cut
tion (Wood and Hopkins 1989, 1992; Broger et al., 1997). An and filled in a single cycle of base-level change and which is
example of a more recent study that illustrates criterion (3) above connected to a marine shoreline; valleys that are located far
is that of Ardies et al. (2002), who, in their study of the Basal inland with no marine link are considered later. We will also
Quartz unit, recognize tributaries and tributary junction scours, assume that fluvial sediment supply and the rate of transgression
both in seismic and by detailed wireline well correlation. are constant, that waves are more significant than tides in the
It is critical when identifying the extent of the incised-valley coastal zone, and that any estuaries that develop are wave-
system to document the geometry of the sequence boundary, dominated (sensu Dalrymple et al., 1992) , because this is the
both inside and outside of the incised valley. The paleotopogra- situation most commonly documented in ancient incised-valley
phy of the incised-valley network (e.g., tributary orientation or systems. For the sake of completeness, we have explicitly in-
valley width/depth) may allow one to determine the cluded the succeeding highstand systems tract, assuming that
paleodrainage direction as an aid in paleogeographic reconstruc- sediment supply is sufficient, relative to the length of the sea-level
tion. In addition, paleovalley networks are proving to be power- highstand, to allow shoreline progradation following the trans-
198 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

gression. At times of high-frequency, high-amplitude sea-level farther up the valley (i.e., the deposits onlap) as the transgression
changes, such as have occurred during the Pleistocene, this proceeds. Ideally, the fill of the seaward portion of the incised
assumption may not be fulfilled, in which case sea level falls and valley (segment 1, Fig. 19) is characterized by backstepping
the river reincises before the estuary is completely filled. (lowstand to transgressive) fluvial and estuarine deposits, over-
lain by transgressive marine sands and/or shelf muds. The
STRATIGRAPHIC ORGANIZATION OVERVIEW middle reach of the incised valley (segment 2, Fig. 19) consists of
the drowned-valley estuarine complex that existed at the time of
Models for incised valleys that are connected to a marine maximum transgression, overlying a lowstand to transgressive
shoreline are based on an ability to subdivide the valley fill succession of fluvial and estuarine deposits like those in segment
longitudinally (Fig. 19) into three segments (as distinct from the 1. The innermost reach of the incised valley (segment 3, Fig. 19) is
tripartite estuarine facies zonation discussed above). This three- developed headward of the transgressive estuarine limit and
fold subdivision reflects the unique depositional and stratigraphic extends to the point where relative sea-level changes no longer
organization of the fill, which results primarily from lowstand control fluvial style. This segment is characterized by fluvial
erosion, followed by transgressive deposition, and finally high- deposits throughout its depositional history; however, the fluvial
stand progradation. style changes due to systematic variations in the rate of change of
As relative sea level falls, the entire length of the incised valley base level. The effect of base-level change decreases inland until
is characterized by (net) fluvial erosion, which creates the basal eventually climatic, tectonic, and sediment-supply factors be-
sequence boundary and may also leave intermediate, falling- come the dominant controls on the fluvial system.
stage terraces within the valley. When relative sea level starts to In the following sections we present additional detail on the
rise after reaching its lowest level, fluvial deposition begins at the characteristics of each incised-valley segment and then provide a
mouth of the incised-valley system and extends progressively range of representative outcrop and subsurface studies.

A B

FIG. 25.—Examples of diagnostic tidal sedimentary structures; A) tidal rhythmites, B) tidal mud drapes in a cross bed that separates
the cross bed into tidal bundles (from MacEachern and Pemberton, 1994).
ESTUARINE AND INCISED-VALLEY FACIES MODELS 199

Segment 1—Outer Incised Valley example of this is provided by the Quaternary sediments in the
Rhine–Meuse valley (Törnqvist, 1993), and in conceptual form
The outer incised valley (segment 1) extends from the most in Figure 21. Note, however, that the amalgamated channel
seaward extent of valley incision, near the lowstand mouth of deposits at the base of the valley fill cannot be assumed to be
the incised valley, to the point where the shoreline stabilizes at braided, simply because of the absence of overbank deposits;
the beginning of highstand progradation (Fig. 19). As in the they could equally well be meandering-river deposits with
other segments, this reach of the valley initially undergoes negligible preservation of muddy overbank deposits because of
fluvial incision with the lowering of base level. Sediment is the low accommodation.
bypassed to the mouth of the valley, where it is deposited as a The thickness of the fluvial succession, and the extent to which
lowstand delta and/or prograding shoreline. This period is the predicted changes in fluvial style are developed, may be
represented by the sequence boundary, which may be overlain variable along the length of segment 1. The ultimate thickness is
by lowstand fluvial to tidal–fluvial deposits (Fig. 20, profile 1). controlled by the accommodation developed during the rise in
As sea level begins to rise and the lower reaches of the system are sea level (Jervey, 1988), with the major factor being the ratio of the
transgressed, the lower reaches of the incised valley change rate of fluvial-sediment input to the rate of sea-level rise. In the
from being a conduit for fluvially eroded sediment to the site of situation where sea-level rise greatly outpaces fluvial input,
fluvial and (subsequently) estuarine deposition. Fluvial deposi- transgression is rapid and the thickness of the fluvial deposits is
tion, although initiated during the late lowstand, continues less than in the case where abundant sediment input occurs
during the early stages of transgression, with the locus of during a slow rise in sea level. In the special case where sediment
deposition shifting landward as relative sea level rises and the input matches sea-level rise, the fluvial deposits aggrade verti-
shoreline transgresses (Wright and Marriott, 1993; Wescott, cally and the shoreline does not transgress. In all cases, the
1993). The transition from erosion and fluvial bypass to fluvial preserved thickness of the fluvial succession may be affected by
deposition migrates landward as the transgression proceeds. subsequent erosion associated with transgression. While this
Thus, the boundary between the lowstand and transgressive fluvial stacking is best preserved in Segment 3 (discussed below),
systems tracts (i.e., the transgressive surface) may lie within the and documented in Arnott et al. (2000, 2002) and Lukie et al.
fluvial deposits rather than at their top and is diachronous if it (2002) from the Basal Quartz Formation, examples of preserved
is picked at a facies boundary. For this reason, the lowstand fluvial stacking controlled by accommodation in Segment 1 are
systems tract (LST—i.e., those deposits that accumulated before found in the Upper Cretaceous of the Kaiparowits Plateau, Utah
the shoreline begins to migrate landward) within the valley may (Shanley and McCabe, 1991, 1994) and the Mesaverde Group
effectively pinch out landward (Figs. 19, 20), although there (Olsen et al., 1995).
should be at least a thin layer (ca. one channel depth thick) of As the transgression proceeds, the estuarine conditions that
lowstand-age fluvial deposits along the length of the valley, are established in the seaward end of the valley migrate land-
unless they have been removed by later channel erosion. Near ward. In a wave-dominated estuarine setting, the first estuarine
the mouth of the valley, most of the fill may be deposited during deposits over the fluvial sediments are tidally influenced fluvial
lowstand-systems-tract time, but farther up the valley the greater and bayhead-delta (distributary channel, levee, and interdistrib-
part of the fill accumulates during transgressive-systems-tract utary bay) deposits (Fig. 20, profile 1). As transgression contin-
time. ues, central-basin deposits then overlie the bayhead delta across
Within the fluvial succession near the river mouth, the early a flooding surface that may correlate updip to a change in fluvial
deposits accumulate when the rate of creation of accommoda- style. The central-basin deposits in turn are overlain by the
tion is low (i.e., near maximum lowstand time), hence channel estuarine flood-tidal-delta and other barrier deposits (cf. Boyd et
amalgamation is common, leading to the formation of a coarse- al., 1992; Dalrymple et al., 1992). This contact may be gradational
grained succession in which muddy overbank deposits are if it corresponds to the prodelta deposits of the flood tidal delta,
scarce. As base level begins to rise ever more rapidly during the but it is equally likely to coincide with the erosional base of a tidal
TST, the fluvial channels become progressively less amalgam- channel (Boyd and Honig, 1992), with the deepest incision occur-
ated and fine-grained deposits are preserved more commonly ring at the location of the tidal inlet. The erosion surface at the base
(Fig. 21; e.g., Boyd and Diessel, 1994). The fluvial style (i.e., of such channels is referred to as a tidal ravinement surface (Allen
braided, meandering, anastomosed, or straight) within the in- and Posamentier, 1993).
cised valley is dependent on a variety of factors, including As transgression proceeds, the shoreface passes the former
sediment supply, grain size, discharge, valley gradient, and rate location of the estuary. Wave erosion associated with shoreface
of transgression (Schumm, 1977, 1993; Schumm and Ethridge, retreat produces a wave ravinement surface that may truncate the
1994). These variables likely change during the rise in sea level underlying estuarine deposits (Fig. 20, profile 1; e.g., Ashley and
associated with the marine transgression (Gibling, 1991; Wright Sheridan, 1994; Belknap et al., 1994; Kindinger et al., 1994; Tho-
and Marriott, 1993; Törnqvist, 1993). Thus, in the simplest case mas and Anderson, 1994). The depth of erosion depends on a
where all other factors remain constant, the character of the variety of factors, the more important of which are:
lowstand to transgressive fluvial sediments should change ver-
tically as the depositional gradient and capacity of the fluvial 1. The depth of the base of the shoreface: a more intense wave
system decreases as the shoreline approaches. This change climate leads to deeper erosion.
would most likely result in successively younger channels
having finer-grained sands than preceding channels, in part 2. The resistance to erosion of the interfluves: lithified bedrock
because of the seaward decrease in grain size within the river is more resistant to erosion than unconsolidated material,
but also because of deposition of the coarser portions of the and may cause the shoreface to ride up and over the valley
sediment load in more inland areas. This overall upward de- fill.
crease in the grain size of subsequent channels accompanied by
and upward decrease in channel amalgamation, with a change 3. The depth of the valley: shallow valleys may be completely
from higher-energy (such as sandy braided) to lower-energy, removed, whereas more of the fill of deeper valleys escapes
(such as mixed-load meandering) fluvial deposits. An excellent truncation.
200 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

4. The rate of relative sea-level rise: more rapid rates of sea-level Senlac heavy-oil pool in southwestern Saskatchewan (Zaitlin
rise promote more rapid transgression, which reduces the and Shultz, 1984, 1990).
potential for deep truncation.
Segment 3—Inner Incised Valley
In many cases, all but the deepest and most landward parts of
the estuary-mouth sand body are removed. Flood-tidal deltas The innermost segment (segment 3) of the incised-valley
and the bases of the deepest tidal channels, including the tidal system lies landward of the transgressive marine–estuarine limit,
inlet, have the highest preservation potential (e.g., Belknap et al., but it is still influenced by changes in base level associated with
1994, Thomas and Anderson, 1994). The wave ravinement sur- relative sea-level change (Fig. 19). This segment may extend for
face may then be overlain by transgressive shoreface to nearshore tens to hundreds of kilometers above the limit of marine/estua-
sands, which may vary in thickness from almost nothing to many rine influence (Shanley et al., 1992; Schumm, 1993; Levy and
meters in shelf sand banks and ridges (Snedden and Dalrymple, Christie-Blick, 1994). The fill of this segment is entirely fluvial,
1998) that were created by shelf processes. Finally, the valley is with no evidence of tidal action or brackish water. Channels may
capped by open-marine mudstones associated with the succeed- be braided, meandering, anastomosed, and/or straight, depend-
ing highstand. The landward limit of these mudstones is an ing on a variety of factors such as sediment supply, gradient,
indicator of the inner end of segment 1. discharge, and sediment size. However, relative sea-level changes
associated with the lowstand–transgression–highstand cycle pro-
Segment 2—Middle Incised Valley duce predictable variations in the rate of creation of accommoda-
tion through time and may also produce a predictable vertical
Segment 2 lies between the inner end of segment 1 (i.e., the succession of fluvial styles (Fig. 20, profile 5; Fig. 21; Gibling, 1991;
initial highstand shoreline) and the estuarine limit (i.e., the Wright and Marriott, 1993).
landward limit of recorded tidal influence) at the time of maxi- Lowstand fluvial deposits are expected to be relatively thin,
mum flooding (Fig. 19). It therefore corresponds to the area because the fluvial system in these inland locations would have
occupied by the drowned-valley estuary at the end of the been erosional or would have acted mainly as a transport
transgression. In this segment the sequence boundary is over- conduit (a bypass zone) at that time. Late lowstand to early
lain by lowstand to early-transgressive fluvial deposits similar transgressive deposits at the base of the fill may be character-
to those in segment 1. These are in turn overlain by transgressive ized by relatively coarse-grained, amalgamated channel depos-
estuarine facies, but in this segment the nature of the overlying its (Fig. 21). As transgression proceeds, an overall upward-
estuarine succession varies along the length of the segment fining succession of channels should be developed as the gradi-
(Figs. 16, 17; cf. Dalrymple et al., 1992) because the estuarine ent and stream capacity decrease as the backwater zone land-
facies are (ideally) preserved with the spatial distribution they ward of the estuary migrates up the valley. The deposits that
had in the contemporaneous estuary. accumulated during times of rising base level should contain
Near its seaward end, (i.e., beneath the preserved barrier that more isolated, channel-sandstone bodies, interbedded with a
forms the landward margin of any subsequent highstand higher percentage of overbank deposits (e.g., Törnqvist, 1993,
strandplain, assuming as we have throughout this discussion Shanley and McCabe, 1994). Freshwater organic facies (e.g.,
that the coastline is wave dominated) the succession is similar to peat or lacustrine carbonates) might be abundant and the soils
that in segment 1, with fluvial and bayhead-delta sediments less mature and wetter than those associated with the lowstand
overlain by central-basin deposits , which are, in turn, capped by (Cross, 1988; Boyd and Diessel, 1994; Wadsworth et al., 2002).
estuary-mouth-barrier sands. Because open-marine conditions The overlying highstand deposits may be expected to coarsen
do not transgress into this segment, the barrier sediments are upward overall, due to progradation in response to decreasing
overlain by highstand fluvial deposits (Fig. 20, profile 2), unless rates of base-level rise and accommodation creation (Schumm,
sea level falls before the estuary fills completely, in which case the 1993).
estuarine deposits are capped by the next sequence boundary. In In terms of relative length, the three incised-valley segments
the middle portion of segment 2, barrier sands are absent, and identified above may be quite variable. If the transgression has
central-basin deposits coarsen upwards above the maximum been extensive, however, segment 1 is likely to be long and may
flooding surface into progradational, bayhead-delta and fluvial extend for most of the width of the formerly exposed continental
sediments of the succeeding highstand deposits (Fig. 20, profile shelf. The length of segments 2 and 3 is related to the depth of
3) that fill the estuary if the highstand is of sufficient duration. valley incision and the gradient above the highstand shoreline.
At the headward end of segment 2, central-basin sediments are For example, on many old, wide passive margins such as the Gulf
absent, and the bayhead delta is overlain directly by highstand of Mexico, segment 1 is much longer than segments 2 and 3
fluvial deposits (Fig. 20, profile 4). The most landward limit of (Thomas and Anderson, 1994; Blum and Törnqvist, 2000). Over-
the detectable marine influence (i.e., tidal features in fluvial all, segment 2 is likely to be the shortest of the three because it
deposits) is taken as the inner end of segment 2. This point corresponds to the length of the estuary at one point in the sea-
corresponds with the inner end of the estuary as defined by level cycle.
Dalrymple et al. (1992), and is also approximately equivalent to
the “bayline” of Posamentier et al. (1988) and Allen and ANCIENT CASE STUDIES OF
Posamentier (1993). INCISED-VALLEY DEPOSITS
Barrier islands are rarely preserved in incised valleys be-
cause typically they are removed by shoreface ravinement The model for an incised-valley fill described in the preceding
during transgression. However, preservation may be possible paragraphs has been applied to a large number of ancient ex-
at the highstand shoreline as the barrier stabilizes and then amples. Here we review several of these to illustrate typical
evolves into a strandplain, as is just beginning on Galveston examples and to highlight controls on the nature of incised-valley
Island, Texas (e.g., McCubbin, 1982). A potential subsurface deposits that are not discussed elsewhere in this chapter, such as
example of such a preserved barrier sand body is provided by the influence of the overall accommodation regime on the char-
the Lower Cretaceous Lloydminster Member (Mannville Group) acter and stratigraphic organization of such deposits. The petro-
ESTUARINE AND INCISED-VALLEY FACIES MODELS 201

leum-industry applications are highlighted in several of these kilometers (Fig. 26). Repeated transgressive–regressive events
examples. developed a compound valley and terrace geometry (Leighton,
1997), similar to that observed in the modern Colorado River in
Case Study 1: Texas (Blum, 1990, 1994). Individual incised-valley systems are
The Mississippian Morrow Formation: between 50 and 80 feet (15–25 m) thick and 0.5 and 2 miles (0.8–
Fluvial to Fluvial–Estuarine Deposits of Segment 1 3.2 km) wide (Krystinik and Blakeney-DeJarnett, 1994; Krystinik,
1989). The incised-valley systems are cut into marine mudstones
A number of well-documented subsurface examples of seg- and limestone of the preceding highstand and are blanketed by
ment 1 incised-valley deposits exist from the Western Interior similar deposits of the succeeding highstand. Multiple unconfor-
Seaway of North America. One such example is the Carbonifer- mity and exposure surfaces merge onto the interfluves.
ous Morrow Formation of the Anadarko Basin, as described by An individual cycle of fill from segment 1 of the Morrow
Krystinik and Blakeney-DeJarnett (1994), Krystinik (1989), and incised valleys consists, from base to top, of:
Bowen and Weimer (1997) (Figs. 26, 27).
The Morrow Formation has been the target of extensive (1) basal (braided) fluvial deposits composed of coarse–me-
exploration over the last forty years and is characterized by dium-grained cross-bedded sandstones (core porosity to 25%;
several well-documented productive trends. The Morrow For- core permeability 0.1–4 darcys), grading upward into
mation is distributed on the north flank of the Anadarko Basin, in
what was a broad, low-relief shelf subject to glacio-eustatic (2) meandering fluvial (core porosity 20–25%; core permeability
exposure and inundation. During glacio-eustatic lowstands (Fig. 0.1–300 md) and floodplain mudstones and green-waxy paleo-
27B), the shelf was largely exposed and subject to fluvial erosion sols that are overlain by
by drainage networks that fed deltas along the rim of the Anadarko
Basin. During interglacial highs (Fig. 27A) the shelf was inun- (3) estuarine (bayhead delta) sandstones (core porosity 3–12%;
dated, with the deposition of mudstone and carbonate. The core permeability 0.1–2 md) and mudstones displaying tidal
shoreline position is thought to have moved in excess of 145–200 influence and restricted bioturbation, topped by
km (90–125 miles) per cycle.
The Morrow incised valleys are characterized by multiple (4) glauconitic sandstone and transgressive marine mudstones
exposure surfaces and fluvial incision interpreted to have been that rest on a shell-rich pebble lag (i.e., the wave ravinement
cut by repeated high-frequency glacio-eustatic sea-level drops, surface) that indicates the Segment-1 character of this ex-
and backfilled with fluvial and fluvial–estuarine deposits during ample. The “hour glass” shape of the well logs through the fill
transgression. The Sorrento–Mt. Pearl–Siaana and Stateline trend (i.e., a basal blocky to fining-upward fluvial to estuarine
is an example of a well-documented Mississippian Morrow struc- succession, overlain by coarsening–upward central-basin to
turally controlled valley that is mappable over hundreds of estuary-mouth deposits) appears to be characteristic in most

FIG. 26.—Map showing the distribution of the Mississippian Morrow Formation incised-valley fills (from Bowen and Wiemer, 2003).
202 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 27.—Schematic diagram showing the distribution of depositional systems during deposition of the Morrow Formation. A)
During relative highstands of sea level, shorelines rimmed the basin and black muds were deposited on a broad, shallow-marine
shelf. B) During relative lowstands, an extensive series of river valleys were developed in eastern Colorado and western Kansas
that flowed into the Anadarko Basin (from Bowen and Wiemer, 2003).

of these incised-valley systems. Outside of the incised-valley produce from backstepping (LST to TST), transgressed fluvial
networks, the interfluve areas are characterized by extensive and estuarine bayhead-delta and central-basin deposits.
paleosol surfaces. The Countess YY pool (Fig. 28), one of several reservoirs
located beneath Lake Newell and adjacent areas in southern
Case Study 2: Alberta, Canada, is interpreted by Broger et al. (1997), Peijs-van
Lower Cretaceous Glauconitic Sandstone: Hilten et al. (1998), and Zaitlin et al. (1998) to lie within segment
Fluvial to Fluvial–Estuarine Deposits of Segment 1 1 of a wave-dominated incised-valley system. A low-permeabil-
ity Middle Glauconitic channel (Fig. 29) incises into the produc-
The Lower Cretaceous Glauconitic Formation of the Western ing channel and locally forms an updip seal to trap hydrocarbons
Canada Sedimentary Basin is characterized by a network of in the Lower Glauconitic channel.
northwestward-trending, compound, piedmont incised-valley Both the Countess YY and Lathom “A” pools contain a num-
systems that are interpreted to feed lowstand to early transgres- ber of characteristic depositional facies that are stacked in a
sive east–west-trending shorelines to the north. The Glauconitic manner that is consistent with the vertical succession proposed
incised-valley system is mappable for at least 535 km south– for a segment 1 incised-valley system. The base of the valley is
north, from Montana into central Alberta (Wood and Hopkins, overlain by fluvial facies that consist of litharenitic, coarse- to
1989, Sherwin, 1994, Broger et al., 1997, Peijs-van Hilten et al., 1998). medium-grained, large-scale trough and planar-tabular cross-
The fills of the 1–5 km wide valleys exhibit a progressive northward bedded sandstone that overlies erosional surfaces that are usu-
change in character. In the south, the fill consists of lowstand to ally covered by a pebble lag (Figs. 30, 31). This facies has excellent
early transgressive, fluvial to fluvial–estuarine deposits character- reservoir quality (Fig. 29), ranges in thickness from 1 m to more
ized by multiple erosive events (i.e., they are compound valley than 10 m, and is encountered at the base of the incised-valley
fills) as a result of low accommodation. In the north, the accom- system. The gamma-ray log signature shows a blocky or fining-
modation was greater and the individual valleys are separated by upward profile. The sediments are interpreted to be deposited by
coarsening-upward highstand shoreface parasequences, result- a highly connected braided to coarse-grained meandering fluvial
ing in full preservation of individual, simple valley fills. channel system.
The Countess–Alderson trend is a 56 mile (90 km) reach of one The bay-head delta facies has moderate to poor reservoir
such Glauconitic IVS that extends over 300 miles (480 km) from quality and either gradationally overlies the fluvial facies or
northern Montana into central Alberta, Canada. Along this reach immediately overlies the basal sequence boundary in areas off
there are 122 hydrocarbon pools (e.g., Countess YY and Lathom the axis of the valley. Thickness ranges from 3 m to more than 11
A pools) that have produced over 100 MMBBL of oil and 300 BCF m. The gamma-ray log signature shows an overall coarsening-
of gas since the 1950s. Recent optimization of many pools using upward trend, indicating a progradational environment. Indi-
a multidisciplinary approach has led to a better understanding of vidual blocky to fining-upward units 3–7 m thick are interpreted
the nature of this incised-valley system. The majority of pools to represent bayhead-delta distributary channels. Single and
ESTUARINE AND INCISED-VALLEY FACIES MODELS 203

FIG. 28.—An example of a compound incised-valley fill from the type Lower Cretaceous Glauconitic Formation well from the Latham
“A” field in southern Alberta, Canada. (Zaitlin et al., 1998). Bottom of core to the lower left; top to the top right. Glauconitic
Sandstone Member; 30, 40, and 50 represent informal units in the Glauconitic; A = fluvial facies, B = tidal–fluvial, C/D = tidal–
fluvial bayhead-delta to central-basin facies.

stacked channel units show a fining-upward and an obvious 1998). Locally an abundance of wave-generated physical sedi-
shaling-upward trend in core and on wireline logs, with tidal mentary structures are present, such as current-ripple lamina-
mud drapes (Fig. 32) being more abundant in the upper parts of tion. Fine mud laminae are present in some intervals. The mud-
the units, indicating either an increase in tidal influence or a stone beds are locally highly carbonaceous, and typically much
decrease in energy during deposition. These bayhead-delta dis- thinner than the intervening sandstone beds (1–5 cm thick). They
tributary-channel deposits are composed of medium- to coarse- commonly contain convolute lamination, syneresis cracks, and
grained, planar-tabular cross-bedded, flaser-bedded, and tidally small-scale gravity faults. Bioturbation is rare in the sandier
bedded sandstones (Fig. 32). Massive to repetitive fining-upward portions of the facies but increases in the mudstone interbeds. The
successions are characterized by basal scour surfaces marked by trace-fossil assemblage is restricted in diversity (Planolites,
shale rip-up clasts and channel lags. From seismic-amplitude Teichichnus, Cylindrichnus, Skolithos, and Tigillites sp.), indicating
maps, a northwest (downvalley) bifurcation of the channel facies the presence of a stress, most likely because of salinity fluctua-
is observed, indicating a distributary-channel pattern (Fig. 33). tions or water turbidity. The heterolithic character indicates
Evidence of tidal activity is indicated by the presence of mud repeated fluctuations in the energy regime, and the sedimentary
drapes and couplets, as well as by a typical estuarine ichnofossil structures indicate that the sand beds were emplaced by density
assemblage in the associated central-basin facies (Fig. 34). flows that were caused by wave, storm, and/or river-flood pro-
In some cases, these channel deposits display a more cesses. The deformation features indicate a depositional slope,
heterolithic character and are interpreted to consist of an inclined and failure of the heterolithic succession. These deposits have
heterolithic tidal point-bar facies that is characterized by a sharp to thicknesses ranging from 8 to 24 m. The gamma-ray log clearly
erosional basal contact with a fining-upward trend. These units shows an irregular alternation of clean sandstone and shale
consist of fine, massive to tidally bedded, flaser-bedded sand- intervals. Reservoir properties and thicknesses of the sandstone
stones, alternating with 2-cm-thick continuous mudstones. All of intervals increase upward, suggesting a sanding-upward and
the strata display a consistent dip of 5–19° and can be considered coarsening-upward trend that indicates progradation. This facies
to be inclined heterolithic stratification (IHS; cf. Thomas et al., is interpreted to be deposited in a bayhead delta-front turbidite
1987) of point-bar origin. Locally, a restricted trace-fossil assem- environment, and is inferred to have a lobate geometry. In some
blage may be present. This facies has moderate to good reservoir wells, the successions show overall lower porosity and perme-
quality and overlies the bayhead-delta distributary-channel fa- ability values than elsewhere, indicating an areal variation in
cies. Thickness ranges from 5 to 7 m. The gamma-ray log signa- grain size or sand proportion.
ture shows an irregular but clearly fining-upward profile. The The central-basin facies consists of fine-grained, rippled, flaser-
abundance of inclined shale intervals increases toward the top of bedded and tidally bedded sandstones displaying abundant
the succession. Sandstone intervals containing tidal mud drapes shale laminae and double mud drapes with a low-diversity
also are more abundant in the top part of this facies. ichnofossil assemblage (Fig. 34). This facies has poor reservoir
The delta-front turbidite facies lies adjacent to the distributary- quality and occurs in intervals with a thickness of 1 to 5 m at
channel deposits and consists of a regular interbedding of planar various stratigraphic positions, most commonly on top of bayhead-
to wavy parallel-laminated sandstones and weakly burrowed, delta sandstone facies and below capping marine shales or the
dark-gray mudstones (Broger et al., 1997; Peijs-van Hilten et al., crosscutting Middle Glauconitic channel sediments. The gamma-
204 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 29.—Porosity (%) versus permeability (millidarcys) cross plot for the various incised-valley facies in the Lower Cretaceous
Glauconitic Formation, southern Alberta, Canada. The data points cluster about four major categories: (1) best reservoir
produceability occurs in fluvial, bayhead-delta channel, and bayhead-delta tidal point-bar (IHS) facies; (2) moderate-reservoir
deposits consisting of sandy, central-basin bay-fill deposits; (3) moderate- to poor-reservoir deposits consisting of muddy,
bayhead-delta fresh-water and central-basin deposits; and (4) Middle Glauconitic lithic-channel facies that locally forms a lateral
seal to the reservoirs because of extensive diagenetic alteration (Broger et al., 1997). BHD = bayhead delta; IHS = inclined
heterolithic stratification; CH = channel; FW = fresh water.

FIG. 30.—A typical basal pebbly fluvial-lag facies from the Lathom “A” 7-19-20-17W4 well, with its associated grain size, porosity (Ø),
and permeability (K) values (from Zaitlin et al., 1998).
ESTUARINE AND INCISED-VALLEY FACIES MODELS 205

FIG. 31.—Sandy cross-bedded fluvial facies from the Lathom “A” 7-19-20-17W4 well, with its associated grain size, porosity (Ø), and
permeability (K) values (Zaitlin et al., 1998).

FIG. 32.—Typical tidal–fluvial facies with tidal couplets and mud drapes from the Lathom “A” 7-19-20-17W4 cored well, with its
associated grain size, porosity (Ø), and permeability (K) values (from Zaitlin et al., 1998). The mud layers reduce permeability
and make this facies a poorer reservoir than the fluvial facies (Figs. 30, 31) that underlie these deposits (see Fig. 29).
206 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 33.—Map of the Lake Newell area, southern Alberta, Canada, showing the distribution of seismic amplitudes along the trend
of the lower Glauconitic incised valley (margins indicated by heavy white lines). Warm colors are interpreted to be either porous
sand (i.e., potential hydrocarbon reservoirs) or undisturbed Ostracod shale, whereas cooler colors indicate nonporous shales.
Note the northwestward bifurcation of the inferred sandstones in the northern part of Lake Newell. This pattern is interpreted
to represent a bayhead delta. The locations of the drillsites are also shown. (Broger et al., 1997.)

ray log signature is irregular. This facies is interpreted to have Case Study 3:
been deposited in the central-basin environment. As is common Lower Cretaceous Senlac (Lloydminster Formation) Sandstone:
in segment 1, no preserved barrier exists in the study area because An Example of an Estuary-Mouth Barrier
it was removed by ravinement. This is particularly true of low- Sandbody of Segment 2
accommodation settings; in areas with higher accommodation,
portions of the barrier (dominantly the tidal inlet and flood-tidal The Senlac heavy-oil pool, located in Townships 38–39, Range
channels that cut down into the central-basin deposit) may escape 26–27W3, of Saskatchewan (Fig. 35), was discovered in 1980. It
removal—e.g., Cretaceous Viking Formation in the Crystal Field has been estimated to contain 1.3 x 107 m3 (84.3 x 106 barrels) of 13–
(Reinson et al., 1988; Pattison, 1991 or in outcrops of the Paddy– 15 degree API oil in a barrier and tidal-inlet complex at the mouth
Cadotte interval (Leckie and Singh, 1991), both in Alberta. of a paleovalley system (Fig. 36). The existence of an intact barrier
ESTUARINE AND INCISED-VALLEY FACIES MODELS 207

FIG. 34.—An example of bioturbated central-basin facies from the Lathom “A” 7-19-20-17W4 cored well, with its associated grain size,
porosity (Ø), and permeability (K) values (from Zaitlin et al., 1998).

complex with its associated estuary is the reason for assigning geneity, variation in lateral continuity, and porosity–perme-
this deposit to segment 2 of the incised-valley model. ability differences associated with original textural characteris-
Four main environments can be identified beneath and within tics. The flood-tidal delta, with increased bioturbation and a
the barrier at the mouth of the paleovalley (Figs. 36, 37): higher proportion of introduced mud, has the poorest produc-
tion characteristics, whereas the tidal inlets have the coarsest
(1) A basal fluvial sandstone to siltstone ~ 5 m thick, organized grain size and the highest initial porosity and permeability,
into repetitive fining-upward cycles of massive to cross- which leads to the most rapid production. Another example of
bedded to rippled sandstone with local rootlets and a re- a preserved segment 2 barrier has been documented in outcrop
stricted trace-fossil assemblage consisting of Paleophycus sections from the Paddy Member of the Albian Peace River
herberti, Conichnus sp., Lokeia sp., and small Thalassinoides sp. Formation (Leckie et al., 1990).
The sands display excellent reservoir quality but are wet,
whereas the siltstone has an effective permeability of < 0.01 Case Study 4:
md and porosity of < 5%. Lower Cretaceous Basal Quartz Sandstone:
A Low-Accommodation Compound Incised-Valley Deposit
(2) An ~ 4-m-thick coal and carbonaceous shale that accumu-
lated in marsh environments. One of the most complex successions of incised-valley de-
posits yet described in detail is provided by the Lower Creta-
(3) A bioturbated central-basin to fringing tidal-flat mudstone. ceous Basal Quartz Formation and its equivalents (i.e., the
Coverley, Lakota, Cutbank, and Sunburst units) in southern
(4) A complex sandbody that consists of upward-coarsening Alberta and northwestern Montana (e.g., Way et al., 1998;
shoreface deposits (effective permeability 2.5 darcys and Dolson and Piombino, 1994; Lukie et al., 2002; Zaitlin et al., 2002;
porosity 27–31%) that are cut by blocky to fining-upward Leckie et al., 2005). The Basal Quartz (BQ) is a relatively thin unit
tidal-inlet channels (permeability ~ 3 darcys and porosity 25– (typically < 100 m) that was deposited in an accommodation-
30%), with back-barrier flood-tidal deltas on its south side limited setting and is characterized by multiple, closely spaced
(permeability ~ 2.7 darcys and porosity 29–31%). unconformities that define a set of more than ten complexly
nested incised-valley fills. The BQ was deposited as part of an
The position and preservation of the barrier imply a wave- elongated NNW–SSE trending foreland trough in which there is
dominated shoreline deposit at the transgressive limit of the pronounced isopach thickening toward the northwest. The
shoreline. trough contains three major north-south paleodrainage systems
There is marked variation in the production history (Fig. 38) (the Spirit River, Edmonton Channel, and McMurray valleys;
between the subfacies of the barrier because of internal hetero- Fig. 39). The older and more southerly occurrences provide
208 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 35.—Two-way travel time in seconds to the sub-Cretaceous unconformity in the Senlac area of southwestern Saskatchewan,
Canada. Darker colors (longer times) indicate areas where the unconformity is deeper. The pattern is interpreted to represent an
incised-valley network. Arrows indicate interpreted paleovalley trends and inferred paleodrainage directions. (From Zaitlin and
Shultz, 1990.)

FIG. 36.—Distribution of inferred depositional environments during Lloydminster Formation time in the Senlac incised valley. SF.
= shoreface deposits; TC. = tidal-channel deposits; FTD = flood-tidal-delta deposits. Heavy black lines separate depositional
environments within the estuary-mouth sand plug. (From Zaitlin and Shultz, 1990.)
ESTUARINE AND INCISED-VALLEY FACIES MODELS 209

FIG. 37.—Idealized vertical sequence of the Lower Mannville Group in the Senlac area. Ichnofossils identified by Dr. G. Pemberton
(University of Alberta); micropaleontological data provided by Robertson Research and Dr. C. Vervoloet. A, B, and C refer to
zones in Figure 36: SB = sequence boundary; IFS = initial flooding surface; TR = tidal ravinement surface; WR = wave ravinement
surface. (From Zaitlin and Shultz, 1990.)

well-documented examples of segment 3 fluvial deposits that divided into four informal mappable units (A Sandstone, Horse-
pass northward into segment 1 fluvial, estuarine, and marine fly, BAT, Ellerslie), each of which can be further subdivided
deposits. (Zaitlin et al., 2002) (Fig. 40). In particular, the A Sandstone has
Within the study area of Zaitlin et al. (2002) (Fig. 39), accom- been divided into the Regional A (oldest), Carmangay, Mesa IV,
modation ranges between the following two end members: and Valley and Terrace units. This informal stratigraphic break-
down was later substantiated by chemostratigraphic analysis of
(1) An area of extremely low accommodation in the southeast the succession (Figs. 42, 43; Ratcliffe et al., 2004). There are two
corner of Alberta, where isopach values range between 0 and cycles of increasing-upward mineralogical and textural matu-
40 m and net sedimentation rates are less than 2.2 m/My. This rity, the first associated with the A Sandstone and the second
area was dominated by long periods of erosion and exposure, associated with the Horsefly–BAT–Ellerslie succession. The
the development of paleosols, and polycyclic incision of subdivision of the BQ into discrete valley systems allows recog-
valley systems characterized by thin, sheet-like, braided to nition of how the paleodrainage changed through time. There is
coarse-grained meandering-fluvial deposits. both a progressive spatial and stratigraphic change in valley
organization, from thin and wide valley forms in the south and
(2) An area of low–intermediate accommodation in the north- at the base of the maturity cycles, to thicker, narrower, and more
west where thicknesses range between 40 m and more than deeply cut systems toward the northwest and top of the cycles
200 m and net sedimentation rates ranged between 1.3 and (Figs. 40, 41). There is also a spatial and temporal change in the
11.1 m/My, and valley systems are less amalgamated and development of tributary systems for the Horsefly–BAT–Ellerslie
more easily mappable, with sheet-like fluvial to coarse-grained (upper) cycle. The Horsefly Sandstone has few well-developed
meandering deposits, paleosols, and thin coals at their bases, tributaries, whereas the BAT is characterized by narrow and
changing upward into finer-grained meandering-fluvial to thin tributaries south of the Vulcan Low, deeply cut complex
fluvial–estuarine systems. tributary patterns within the Vulcan Low, and linear deep
tributaries north of the Vulcan Low (Ardies et al., 2002; Zaitlin
The transition between these two areas corresponds closely to et al., 2002).
a geophysically defined ENE-trending structural zone termed The style of depositional fill also changes stratigraphically
the Vulcan Aeromagnetic Low (Ross et al., 1997). and spatially, from braided and coarse-grained meandering-
The BQ has an extensive data base of wireline logs, cores, fluvial sheet deposits in the Regional A Sandstone, Carmangay,
cuttings, and producing pools that allow the succession to be and Horsefly units south of the Vulcan Low and in the low-
210 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 38.—Plot of cumulative oil production vs. time for three typical wells completed in the tidal-channel, shoreface and tidal-delta
lithofacies. Inset: total cumulative oil production vs. time for the Senlac Pool. (From Zaitlin and Shultz, 1990.)

accommodation portions of the Valley and Terrace and BAT units grained cross-bedded to rippled, well sorted sandstones. These
north of the Vulcan Low, to meandering-fluvial deposits associ- sandstones may be capped by thin, variegated to green waxy
ated with somewhat higher-accommodation Mesa IV, Valley and paleosols that formed during long periods of exposure. Partial
Terrace, Horsefly, and BAT units, and then to fluvial–estuarine pedogenic clay plugging is pervasive and typically degrades the
deposits in the portions of the Valley and Terrace, BAT, and porosity and permeability of the Mesa IV deposits. The Mesa IV
Ellerslie units, which accumulated in the highest-accommoda- valleys contain narrow, sinuous, ribbon-like channel deposits,
tion settings north of the Vulcan Low (Figs. 40, 41). less than 15 m thick and 1.6 km wide. Individual channels are
The Carmangay unit (Figs. 40, 41) forms a thin sheet-like difficult to map in the absence of core. The Mesa IV deposits are
sandbody, up to 20 m thick, in the southwest corner of the study interpreted to have been formed by coarse-grained meandering-
area and is interpreted to have accumulated entirely in segment fluvial systems. Locally, the Mesa IV sandstones constitute fair to
3. It consists of multiple cycles of erosionally based, fining- excellent gas reservoirs with 12–25% porosity and < 0.1 md to 0.8
upward channel deposits, 1–5 m thick, of medium- to coarse- darcy permeability.
grained, pebbly, cross-bedded sandstones, fining upward into The Valley and Terrace deposits (Figs. 40, 41; cf. Hamilton et
fine- to medium-grained cross-bedded to rippled, well sorted al., 2001) consist of braided and coarse-grained meandering-
sandstones. Where preserved, the cycles are locally capped by fluvial deposits, grading upward into fluvial and interbedded
thin variegated to green waxy paleosols. During Carmangay floodplain and paleosols. Toward the north the Valley and Ter-
time, braided-fluvial to coarse-grained meandering-fluvial sys- race deposits contain tidal–fluvial channel and estuarine central-
tems migrated across the depositional surface. The lateral migra- basin deposits of segments 1 and 2. As the name indicates, the
tion of the channels effectively removed most fine-grained Valley and Terrace unit consists of a series of nested terraces that
overbank deposits and left multiple basal scour surfaces. Reser- formed during repeated periods of base-level fall and subsequent
voir parameters range from < 0.01 md to 4 darcy permeability, backfilling at a time of overall falling base level, resulting in an
with 20% porosity, that yield excellent reservoir quality. architecture that is similar in style to the Quaternary Colorado
The Mesa IV unit (Figs. 40, 41) also lies entirely in segment 3 River (Blum, 1990, 1994; Blum and Valastro, 1994; Blum et al.,
and consists of multiple cycles of erosionally based, fining- 1994). Southward-directed transgression of the northern Boreal
upward medium- to coarse-grained pebbly cross-bedded quartz Seaway during Valley and Terrace time resulted in the backstep-
and rusty-chert sandstones fining upward into fine- to medium- ping of estuarine deposits over fluvial deposits. Reservoir param-
ESTUARINE AND INCISED-VALLEY FACIES MODELS 211

FIG. 39.—Isopach map of the Lower Mannville Group in the Western Canada Sedimentary Basin. The northwestward increase in
thickness indicates that accommodation increased in that direction during deposition of the Basal Quartz. L = low-accommoda-
tion area; I = intermediate-accommodation area; H = high-accommodation area. Arrows indicate paleodrainages of the
McMurray, Edmonton, and Spirit River valley systems. BC = British Columbia; AB = Alberta; SK = Saskatchewan; MB = Manitoba;
MT = Montana. (From Zaitlin et al., 2002.)

eters range from 5–28% porosity, and 0.06 md to 1.2 darcy The channel deposits in the Horsefly Sandstone exhibit a
permeability. classic upward change from amalgamated to isolated (cf. Fig. 21).
The Horsefly unit is confined to two major compound incised- The cycle begins with a regionally mappable erosional surface
valley systems termed the Whitlash Valley (Hayes, 1986; Hayes that is overlain by amalgamated braided-fluvial sandstones. Any
et al., 1994) and Taber–Cutbank Valley (Lukie, 1999; Lukie et al., contemporaneous overbank mudstones were completely eroded.
2002; Arnott et al., 2000, 2002), both of which extend southward These sandstones are then overlain by mudstone-dominated
into northern Montana (e.g., Dolson and Piombino, 1994), where overbank deposits that encase “ribbon” channel and sheet-like
the Horsefly is termed the Cutbank Sandstone. The Horsefly crevasse-splay deposits (Arnott et al., 2000, 2002; Lukie et al.,
succession is up to 25 m thick, and the Taber–Cutbank Valley is 2002; Zaitlin et al., 2002). Two such successions are present within
approximately 50 km wide (Fig. 41). The valley fill consists of the Horsefly. Each of these sequences accumulated under condi-
repeated fining-upward successions of braided-fluvial to coarse- tions of continuously increasing accommodation. Tectonic move-
grained meander sandstones overlain by thick successions of ments, perhaps in response to episodic thrust loading, are thought
muddy paleosols. The basal strata consist of poorly sorted, ma- to have been the major control on accommodation; eustatic
trix-supported conglomerate with a medium- to coarse-grained fluctuations were probably not important because the study area
sandstone matrix. Clasts are subrounded and several decimeters lay far inland at the time of deposition, landward of the landward
in diameter and are composed of sandstone and silty mudstone. limit of estuarine conditions (i.e., in segment 3). Reservoir param-
The basal unit is overlain by cross-stratified upper medium- to eters of the Horsefly unit range from 3–24% porosity and < 0.01
coarse-grained sandstone, gradationally overlain by massive to md to > 1.2 darcy permeability.
small-scale cross-stratified fine-grained sandstone, in turn over- The BAT can be divided spatially, on the basis of deposi-
lain by siltstone and silty mudstone. The overlying paleosol tional style, into two sub-units (Fig. 41; Zaitlin et al., 2002). The
deposits are composed of variegated red, green, and gray silt- first is a low-accommodation BAT in areas where the total BAT
stones and mudstones that locally reach 30 m in thickness (Lukie, unit isopach is less than 30 m. South of Township 20 along the
1999; Lukie et al., 2002; Arnott et al., 2000, 2002; Zaitlin et al., 2002). Taber–Cutbank valley system the width is of the order of 1–5
212 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 40.—Variations in BQ valley form and width:depth ratios. Cycles 1 and 2 relate to tectonic stages of the adjacent Cordillera during
accumulation of the Basal Quartz (as defined in Zaitlin et al., 2002). Note how in each cycle the first valleys are broad and relatively
shallow, whereas younger valleys have greater depth-to-width ratios.

km. The second is a high-accommodation BAT along the natural that there have been suggestions that the model fails to
Carseland–Crossfield–Penhold trend, and the Provost trend take into account important variables and thus does not accu-
(where the term Dina is used), where total isopach values can rately reflect certain aspects of estuarine and incised-valley de-
reach up to 100 m and the valley width is approximately 6 to 10 posits. There have also been attempts to develop refinements
km and the valley-filling deposits are characterized by fluvial– and/or elaborations of the model, in the same way that the
estuarine deposits of segment 1. In low-accommodation areas, models for a meandering-river point bar have multiplied from
the BAT consists of stacked, erosionally based, fining-upward the single vertical succession proposed by Allen (1963) to the 16
sheet-like sandstones (Ardies, 1999; Ardies et al., 2002; Arnott et successions shown by Miall (1996). In addition, there has been
al., 2000, 2002; Zaitlin et al., 2002). Each succession grades inadvertent misuse of the model by some workers. Here we
upward from coarse to medium sandstone, to lower medium– examine some of the issues raised by these developments, be-
upper fine sandstone. All of these sandstones are pervasively cause they illustrate useful information about incised-valley es-
cross stratified and are interpreted to have accumulated in tuarine systems or about the nature of facies models in general.
braided to coarse-grained meandering rivers. Very rarely does
the low-accommodation BAT display any form of marine bio- Estuary Versus Estuary:
turbation. The BAT sandstones display excellent reservoir qual- The Implications of Applying a Name
ity and are a prime exploration target. In low-accommodation
BAT reservoirs in the southern and eastern portions of the study One of the most fundamental problems with estuarine facies
area, reservoir parameters range from 3–28% porosity and < models has been the ongoing confusion between the oceano-
0.01 md to 5 darcy permeability. graphic, salinity-based definition of estuaries (Pritchard, 1967)
and the modified geologic definition of Dalrymple et al. (1992)
CRITICISMS, MISUSES, AND REFINEMENTS used here. This, in turn, has led to the potential for inaccurate
OF THE E&IV MODEL interpretations of ancient successions and/or to suggestions that
one or other of the definitions is inappropriate. At the outset, it
The E&IV facies model detailed above (Dalrymple et al., 1992; must be recognized that both definitions are “valid” in their own
Zaitlin et al., 1994) has gained widespread usage (Fig. 3) and right. The problem arises through failure to carefully articulate
acceptance over the past decade and could now be regarded as a which definition is being used and/or to implicitly switch be-
mature and established model. However, like all facies models tween definitions without saying so.
that are necessarily based on a “distillation” of natural variability The most common expression of this problem is the growing
(Walker, 1984b), it does represent a simplification of natural tendency to deduce that certain ancient deposits accumulated in
complexity and cannot be expected to match every specific ex- an area of brackish water, on the basis of the nature of the trace-
ample, whether modern or ancient. As a result, it is perhaps fossil assemblage as described below. From this, the authors state
ESTUARINE AND INCISED-VALLEY FACIES MODELS

FIG. 41.—Cycle 1 (lower right): Composite isopach and paleogeographic map of the A Sandstone. Contour interval = 5 m, values range from 0 to 30 m. Cycle 2 (lower left)
Isopach and paleogeographic map for the Horsefly unit. Cycle 2 (upper left): Isopach and paleogeographic map of the BAT unit. Cycle 2 (upper right): Isopach and
paleogeographic map of the Ellerslie unit. In all panels, arrows represent inferred paleodrainage direction. Where no contours are present, no deposits are present,
due either to nondeposition or to subsequent erosion. Original from Zaitlin et al. (2002).
213
214 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 42.—Vertical changes in the geochemistry of silty claystone in the Horsefly, BAT, and Ellerslie units of the Basal Quartz Formation
(Zaitlin et al., 2002; Ratcliffe et al., 2004). The data come from several cores, with samples placed in their correct, relative
stratigraphic position. Al2O3 and SiO2 values demonstrate that there are only minor differences in the silt and clay content of the
various units. However, to minimize the influences of subtle changes in silt content, the values for the other elements have been
normalized against Al2O3.

FIG. 43.—Cross plots of normalized elemental ratios to illustrate differentiation of the Horsefly, BAT, and Ellerslie units (Basal Quartz
Formation) using geochemical data.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 215

that the deposits are “estuarine”, implicitly utilizing the salinity- added to the two-fold wave- and tide-dominated subdivision
based definition of Pritchard (1967). Then, on the basis of this proposed by Dalrymple et al. (1992). Such a proposal would seem
estuarine interpretation, the deposits are said to be transgressive reasonable by analogy with the three-fold subdivision of delta
and/or to demonstrate the existence of an incised valley, which facies models (Coleman and Wright, 1975; Galloway, 1975). It is
implicitly represents a switch to the Dalrymple et al. (1992) certainly the case that there is a wide range in the size of rivers
geological definition. Alternatively, the authors might demon- feeding estuaries. However, this proposal for a river-dominated
strate that the succession is, in fact, regressive and go on to class of estuary has weaknesses for three reasons:
suggest that progradational estuaries exist, in contravention of
the Dalrymple et al. (1992) definition. (1) In the specific instance described by Cooper (1988), the short-
Such switching between the two definitions of estuary is term and long-term behavior of the system was not ad-
inappropriate because, as stated above, the salinity-based defini- equately taken into consideration. Because of the extreme
tion includes a much broader range of environments than the variability of discharge in that situation, the river-mouth area
geological definition: although the two definitions overlap in alternated between two conditions: at the time of the infre-
their application, they are not equivalent. It is certainly the case quent but very large river floods, sand was exported beyond
that estuaries (sensu Dalrymple et al., 1992) may have a phase of the mouth of the river to the marine environment, whereas,
progradational filling at the end of the transgression, when the during the much longer, intervening periods, the river-mouth
coastal zone switches from transgression to regression. However, area was refilled by sand carried to the area by river and
this progradational phase must overlie a transgressive succes- flood-tidal processes. During the times when sediment was
sion. Furthermore, the application of the term “estuary” sensu being imported, the system was a wave-dominated estuary
Dalrymple et al. (1992), in combination with the (now modified) (sensu Dalrymple et al., 1992) with a barrier, flood-tidal delta,
idea that such estuaries are restricted to incised valleys, also and central basin). In the longer term, however, sediment was
implies that there was a relative sea-level lowstand and the being supplied by the river to a beach and shoreface system.
development of a sequence boundary at the base of the valley (cf. As a result, in the longer term the system described by Cooper
Hein and Langenberg, 2003). This, in turn, has important impli- (1988) is not an estuary but is a river feeding an incipient
cations for our understanding of the geological history of the area strandplain.
and for the prediction of petroleum-reservoir play types (e.g.,
lowstand deltas). However, brackish-water trace-fossil assem- (2) In the more general sense, one of the most important, even
blages can occur in progradational deltaic settings and even in defining, characteristics of estuaries (sensu Dalrymple et al.,
some shelf environments. It may be, therefore, that the inappro- 1992) is the existence of two sediment sources: fluvial and
priate switching between the two definitions of estuary has led to marine. In the limiting cases where one or other of these two
the misidentification of deltaic distributaries as estuaries and the sediment sources goes to zero, it is legitimate to argue that the
incorrect sequence-stratigraphic interpretation of some succes- systems are no longer estuaries in the original sense. Therefore,
sions (cf. Reinson and Meloche, 2002; Zaitlin, 2003; Krystinik and systems with only a marine sediment source and no river
Leckie, 2005). influence might legitimately be considered barrier–lagoon
It should be noted that the definition of estuary presented in systems that are gradational with estuaries (cf. Boyd et al.,
this paper is a modification of that presented in Dalrymple et al. 1992). Such systems, in our opinion, form exclusively in trans-
(1992). Since the original definition was constructed it has become gressive situations. Systems with negligible marine sediment
apparent that there are numerous settings such as abandoned input (i.e., they are “river-dominated”) are, by contrast, almost
deltas and structural embayments that possess the characteristics certainly regressive, at least locally at the river mouth, at the
of estuaries but are not necessarily associated with paleovalleys. time of consideration. Therefore, they fail to fulfill one of the
While the origin and classification of these types of settings are fundamental criteria of “estuary” (sensu Dalrymple et al.,
usually apparent in modern environments, it is much more 1992). There is no good, existing term for a semi-enclosed
difficult to discern them in ancient sediments. Hence, while it coastal area with no marine input that might otherwise be
might be preferable to identify an abandoned delta as such in an called river-dominated. One possibility would be to call such
ancient deposit, it might not always be possible to do so, in which systems “embayments”, as is commonly done in the coastal
case the use of the term estuary would be justified if it met the geomorphological literature (e.g., an open-mouthed bay with
criteria identified in this paper. no bay-mouth barrier or other marine-sourced sediment body,
The types of problems that may result from switching be- but with river input at its head). Therefore, given the essential
tween the two definitions of estuary represent inappropriate use character of estuaries as proposed by Dalrymple et al. (1992) a
of the models rather than deficiencies in the definitions. Both prograding river-dominated system cannot be an “estuary”.
definitions of estuary have their use, but they should not be
confused. We suggest that, if the Dalrymple et al. (1992) definition (3) A careful review of modern river-mouth areas (cf. Dalrymple
is to be used at any point in a study, the salinity-based definition et al., 1992) indicates that the size of the river does not
be avoided. Instead, we recommend the use of the term “brack- fundamentally change the geomorphic character of the estua-
ish-water” as the more acceptable term (in place of “estuarine”) rine system. Therefore, valley mouths that have unfilled
for deposits believed to have accumulated in an area of reduced accommodation (i.e., they are estuaries sensu Dalrymple et
salinity. Conversely, if the decision is to use the salinity-based al., 1992) have similar morphologies regardless of whether
definition, then the Dalrymple et al. (1992) definition should be the river is small or large. For example, the Severn River
avoided and the term “transgressive” should be used for (England) and Salmon River (Cobequid Bay, Bay of Fundy)
retrogradationally stacked facies successions. tide-dominated estuaries have essentially identical morpho-
logical and facies zonations despite the fact that the water and
Classification of Estuaries sediment discharges of the Severn River are several orders of
magnitude larger than those of the Salmon River. Similarly,
Several authors, beginning with Cooper (1988), have sug- the fundamental morphology of the large Mobile Bay estuary
gested that a third type of estuary (fluvially dominated) should be (Kindinger et al., 1994) is identical to that of the small
216 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

Narrawallee and Wapengo estuaries of southern New South the nature of the valley fill in detail and show tens of kilometers
Wales (with coast-parallel barrier, low-energy and muddy of backstepping of facies, which clearly lie within the TST as
central basin and bayhead delta; Nichol, 1991). defined by most workers, but which they say forms part of the
LST. Such inconsistent use of terminology is confusing at best and
Thus, the creation of a river-dominated class of estuaries deviates from the original intent of systems tracts.
would appear, at least at this time, to be unnecessary. People
working in the ancient rock record who have adopted this con- Relative Abundance of Facies and
cept may have fallen victim to the inadvertent mixing of estuary Systems Tracts within Incised Valleys
definitions discussed in the preceding section.
The original model for incised valleys (Figs. 19, 20; Zaitlin et
Systems-Tract Assignment of Valley Fills al., 1994) shows fluvial deposits as constituting a very small
proportion of the entire valley fill, which was dominated by
Some confusion exists regarding the assignment of incised- estuarine facies. As a result, the TST was volumetrically predomi-
valley fills to individual systems tracts. The original work on nant, with minimal LST. While these authors explicitly said that
incised-valley deposits by the Exxon group (e.g., Van Wagoner et the relative proportion of fluvial (and LST) deposits was subject
al., 1988, Posamentier and Vail 1988) considered all of the depos- to considerable variability, some subsequent workers have criti-
its within an incised valley to belong to the LST. In this context, cized the model, suggesting that this is not a universal aspect of
this was reasonable because they were dealing with relatively incised-valley successions.
low-resolution seismic data and large-scale stratigraphic se- Such criticisms may have some validity, but they fail to
quences of second or third order. In this context, the fill of the recognize the nature and role of facies models. As already stated,
valley could not be subdivided in detail and the valley-fill succes- facies models represent a distillation of existing knowledge and
sion represented a very small volume at the base of the much are not intended to illustrate the only possible stratigraphic
larger sequence. By contrast, detailed examination of both modern expression. Variability is to be expected, and deviations from the
and ancient valley-fill successions (e.g., Roy, 1984; Reinson, 1992; model can be used to deduce important information about the
Boyd and Honig, 1992; MacEachern and Pemberton, 1994; Demarest situation under study. For example, the complete absence of
and Kraft, 1987) show clearly that a significant fraction of the fluvial deposits and the presence of tidally influenced deposits
valley-filling deposits in many systems was deposited during right to the base of the valley may indicate either (1) that the
base-level rise, commonly at a time when the shoreline had mi- erosional feature is not a valley but instead represents a tidally
grated substantial distances landward of its lowstand location. scoured depression that may not correlate to a sequence bound-
Incised-valley estuaries along modern coastlines illustrate this ary or (2) that deposition within a valley took place near the
point: valley filling continues at a relative highstand of sea level. As lowstand river mouth in a tidal–fluvial environment. On the
a result, many, but not all, workers have tended to recognize both other hand, a valley filled entirely with fluvial deposits indicates
LST and TST deposits within incised valleys, with TST deposits that the rate of fluvial sediment supply was high relative to the
predominating throughout most of the length of the valley. rate of creation of accommodation by sea-level rise, or that the
Although this situation is perhaps the “norm” (sensu Walker, location in question lay sufficiently far inland that estuarine
1992), valleys, or portions of valleys, that are filled entirely during conditions never reached there (i.e., the valley lies within seg-
the lowstand are a possibility. In particular, this may occur for a ment 3). In retrospect, the original Zaitlin et al. (1994) representa-
distance landward of the lowstand shoreline, with the valley fill tion with minimal fluvial and LST deposits may have been
consisting of fluvial deposits that accumulated during the fluvial unduly influenced by the then predominance of systems in which
aggradation that accompanied sea-level rise during the late LST there was a relatively small fluvial sediment input and of modern
and earliest TST. High rates of sediment supply at lowstand systems in which the rate of RSL rise was so rapid that minimal
would favor valley filling at this time. However, the inland extent fluvial–LST deposition occurred, especially in the inner part of
of this lowstand fluvial aggradation would be limited to the area segment 1 and in segment 2 (e.g., incised-valley systems along the
where the “backwater” effect exists (a few kilometers to several US east coast such as described in Ashley and Sheridan, 1994). A
tens of kilometers at most; e.g., Blum and Törnqvist, 2000, and better “distillation” might well have included more fluvial sedi-
references therein) during the late lowstand. At the same time, ment as the “norm”.
areas farther landward in the valley would be bypass zones with
little or no net deposition. As the lowstand shoreline experienced Additional Critiques of Estuarine and Incised-Valley Models
initial transgression at the onset of the RSL rise, the transgressive
surface would be formed. This surface, where it is possible to Other discussions of the E&IV models have been published by
recognize it, would onlap into the valley. Landward of the point Washington and Chisick (1994) and Blum and Törnqvist (2000).
of onlap of the transgressive surface, the valley fill would consist Washington and Chisick (1994) suggested that several factors
of a thin LST (possibly only one channel depth thick in many were missing from the estuary model of Dalrymple et al. (1992).
cases) consisting of relatively coarse-grained fluvial deposits, They identified the in situ production and accumulation of bio-
overlain by finer-grained fluvial sediments of the TST. genic material (peat and carbonate), the rate of sea-level rise
Both systems-tract assignments of valley-fill deposits are relative to the rate of marine sediment input, and climate (tem-
probably valid, but at very different scales of consideration. The perate versus tropical) as factors that should have been included.
early Exxonian view that all valley-fill deposits are LST should be In response we note that no generalized model can include all
used only at very large spatial and temporal scales, whereas a factors that are present in a depositional sedimentary environ-
more refined subdivision into LST and TST is more likely to be ment. The full range of boundary conditions and processes in an
correct in high-resolution studies. In our opinion, an example of environment determines the spectrum of deposits that may be
what can happen by an inappropriate use of the Exxonian view in produced by that environment; however, only the commonly
a high-resolution study is provided by Bowen and Weimer (1997, occurring combinations will be useful for a widely applicable
2003). In these papers, the authors use the Exxonian approach model. Hence, while the three factors identified by Washington
without clearly explaining why. They then proceed to document and Chisick (1994) may be important in local examples, the lack
ESTUARINE AND INCISED-VALLEY FACIES MODELS 217

of explicit inclusion of them in the Dalrymple et al. (1992) model tidally influenced facies should be considered an estuarine de-
illustrates the distillation process identified by Walker (1992), by posit, a suggestion that implicitly follows the Pritchard (1967)
which variability is removed and generalized facies models are definition of an estuary. It was also suggested that many deltaic
produced. In the case of the three factors above, they are not deposits had been incorrectly identified as estuaries because of
included in the general model because: (1) they do not control the the recent popularity of E&IV models (e.g., Reinson and Meloche,
basic geomorphic organization of estuarine facies; (2) their influ- 2002; Leckie and Krystinik, 2005). This illustrates the need for
ence is less pervasive or less intense than that of the fundamental practical and accurate facies models, because brackish-water
interaction of fluvial and marine processes; and/or (3) the nature tidal facies actually occur in several distinct environments, and
or distribution of their influence is controlled by the fluvial– because deltas should not be confused with estuaries. By devel-
marine interaction in an estuary (i.e., the latter factor is more oping clear facies models based on distinctive combinations of
fundamental; cf. Dalrymple et al., 1994a). As estuarine facies sedimentary processes it is possible to correctly identify and
models become progressively more refined, however, future differentiate these environments.
workers might well wish to create a “new” facies model (i.e., a While it was noted in an earlier section of this paper that each
variant on the models proposed by Dalrymple et al., 1992) to facies model necessarily is a simplification of a wide spectrum of
explicitly incorporate the distribution of carbonate facies in tropi- similar environments, there should be fundamental differences
cal estuaries with low fluvial influence. in facies models from different depositional environments. So
More recently, Blum and Törnqvist (2000) have criticized how facies models from estuaries, tide-dominated deltas, lagoons,
some workers have used the incised-valley concept because it and tidally influenced shelves should not be the same. For ex-
implies a “vacuum cleaner” approach to fluvial sediment trans- ample, the differentiation of tide-dominated deltas from tide-
port rather than a “conveyor belt” approach. Blum and Törnqvist dominated estuaries provides a convincing argument for the
(2000) have disputed the influence of relative sea-level fall as the clear establishment of facies models for each setting and their
initiator of incision, accompanied by “an upstream-propagating appropriate use, and it represents an important example of the
wave of stream rejuvenation, which produces sediments that value of the facies-model concept.
entirely bypass the coastal plain and newly emergent shelf to The delta-versus-estuary problem was formally raised by
provide a critical volume of sediment for systems tracts further Walker (1992), who suggested that the triangular classification of
basinward” (the vacuum-cleaner model that results in an incised deltas (e.g., Galloway, 1975) was inappropriate and that it should
valley). This is contrasted with the conveyor-belt model, “where be modified or abandoned. Walker’s (1992) emphasis on sea-level
sediments are continuously delivered to the basin margin from a change and the presence of a coastal protuberance (i.e., a bulge)
large inland drainage”. Instead, they suggest that it is the climati- as a distinguishing feature of deltas led him to believe that tide-
cally produced changes in discharge that drive incision. While dominated deltas, which commonly occur at the heads of
this may be true in many cases and is not explicitly considered in embayments, were not related to other deltas and were better
many discussions of valley formation, it is hard to neglect the role considered as tidal estuaries. However, the problem results from
of relative sea level (RSL) fall as a trigger for valley formation, a fundamental confusion of the factors that make up the essence
because a fall in RSL may cause the river to encounter new areas of facies models for deltas and estuaries.
of steep gradient on the continental shelf that promote incision. Many of the detailed features of tide-dominated deltas and
Although the impact of these new gradients is not felt throughout tide-dominated estuaries are certainly similar. For example,
the drainage basin, incision is present on many rivers 40–400 km both of them contain brackish water and hence restricted
upstream of the present shoreline (data of Blum and Törnqvist, faunal and ichnological assemblages. They contain very simi-
2000). Indeed, the very abundance of Holocene incised valleys lar physical sedimentary structures (e.g., tidal bundles, in-
containing estuaries on many coastal streams near the highstand clined heterolithic stratification, and all other tidal indicators
shoreline points to the strong influence of RSL change on their listed in recognition criteria 8 and 9 above), as well as similar
development. So, although much sediment is transported through depositional sub-environments and facies (e.g., tidal–fluvial
alluvial valleys in response to climatic forcing during times of channels and elongate tidal sand bars). However, that is as far
sea-level fall and steeper shelf gradients, some sediment is also as the similarity goes. There are fundamental differences that
removed from the coastal plain, generating a container (the distinguish the two depositional environments and their fa-
valley) for later filling. This line of argument highlights one of the cies models.
new features to emerge since the development of E&IV models,
namely the recognition that valley incision may take place only at (1) Estuaries are commonly associated with incised valleys while
localized changes in gradient where knickpoints can be created. deltas are not. However, early highstand progradation of
Hence, although full cross-shelf incision may occur when the some deltas may be restricted to incised-valley settings, while
shoreline drops below the shelf break (e.g., Suter et al., 1987; some abandoned deltas that are not incised take on an estua-
Por™bski and Steel, 2003), a more common situation results from rine character during transgression.
sea-level change that exposes a local gradient increase at an old
shoreface. This may cause incision at several localized sites while (2) Estuaries display a tributary pattern (see above) while deltas
the greater part of the exposed continental shelf and the upstream display a distributary pattern.
alluvial channel remain unincised (e.g., Woolfe et al., 1998;
Posamentier, 2001; Fielding et al., 2003; Wellner and Bartek, (3) Deltas have only one sediment source and hence single
2003). composition, while estuaries have two.

AN EXAMPLE OF FACIES-MODEL USAGE: (4) Deltaic sands fine unidirectionally seaward while estuaries
THE TIDE-DOMINATED DELTA VERSUS show a grain-size peak at either end of the system, reflecting
ESTUARY CONTROVERSY the two sediment sources.

At a recent SEPM research conference (Dalrymple, 2003) it (5) Deltas are fundamentally regressive systems while estuaries
was suggested by some participants that any brackish-water, are transgressive.
218 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

(6) Because deltas are regressive in nature, their stratigraphy followed by a need for classification of the observations. Both of
differs fundamentally from transgressive estuaries. In deltas, these stages occur early in the development of a discipline. As the
marine sand bars are underlain by prodelta and marine field advances, however, classification gives way to the develop-
sediments. In estuaries, marine sand bars are underlain by a ment of empirically based laws and finally to theoretical under-
tidal ravinement surface and more landward estuarine and standing (Hempel, 1965).
fluvial facies (Figs. 16, 17, 19, 20). The field of facies models is still a relatively young field with
a history of less than forty years. Hence, we are in the early stages
(7) The prodelta environment is missing in estuaries. of its development, in which we have made a large number of
observations in the form of surveys and process measurements in
(8) The estuary typically lies on a regional unconformity or on modern environments, outcrop studies, wireline-log, core and
fluvial deposits , which in turn lies on an unconformity. It has borehole studies, and remote-sensing studies (e.g., seismic, ra-
a maximum flooding surface located within or above the dar). These observations have been incorporated into deposi-
estuarine fill. A highstand delta typically lies above a maxi- tional facies models since the middle of the twentieth century in
mum flooding surface and has a sequence boundary devel- what is essentially a form of classification. An approach of this
oped above it (Fig. 19). kind describes the delta and coastal classification triangles pre-
sented earlier.
(9) In sequence-stratigraphic terms, estuaries more commonly Inherent in this approach is an organization of the processes
occupy the transgressive systems tract while deltas more that control deposition and hence involves some understanding
commonly occupy the highstand systems tract (although it is of the relationships between the controlling parameters (in these
recognized that these depositional systems can occur in a cases, for example, waves, tides, and rivers). Therefore, our
range of systems tracts, especially when considering lower- scientific field is at the point of transition to the next stage , which
order sequences). involves empirical approaches and finally theoretical approaches
to understanding.
So, while many aspects of tide-dominated estuaries and del- In the E&IV field, empirical laws have been developed and
tas look superficially similar, they should not share the same applied, for example, to paleohydraulics (Miall, 1996), simula-
facies model. When the correct identification of estuarine and tion of alluvial stratigraphy (Bridge and Leeder, 1979), the
deltaic deposits in their appropriate stratigraphic context is made, influence of relative sea level on river incision (Wood et al.,
it is clear from the nine issues listed above that there are funda- 1993), the continent-wide quantitative classification of coastal
mental differences in the two facies models. Our conclusion is systems based on physical processes (Harris et al., 2002) and the
that the “offending corner of the delta triangle” that was removed preservation of estuarine strata after shoreface erosion (Cowell
by Walker (1992, his Figure 7) should be firmly reaffixed. In et al., 1999). All of these examples and many others have begun
addition, it should be placed correctly in the triangular coastal to take a quantitative approach to sedimentation problems with
classification of Boyd et al. (1992) and separated as shown in the ultimate aim of achieving a theoretical understanding. We
Figure 9 from tide-dominated estuaries. The reasoning behind believe that the quantitative approach to sediment modeling is
this return to the triangular classification is the contrasting pro- the best way to advance our field. Our current stage of develop-
cesses that distinguish deltas (e.g., Wright, 1985) from estuaries. ment is the formulation and application of facies models, with
Chief among these is the balance between sediment flux and a resulting proliferation of these models. The way to avoid
relative sea-level rise. In deltas, over a longer term, the sediment becoming bogged down in this classification stage, as also
flux outstrips any change in relative sea level, while in estuaries occurred, for example, in the study of cyclothems (e.g., Wanless
the reverse is true. In deltas, the fluvial processes delivering and Weller, 1932) or geosynclines (e.g., Kay, 1951) in the earlier
sediment to the coastline overwhelm the marine processes be- twentieth century, is to employ a quantitative approach to
cause there is no available onland accommodation, and they determine the predictive relationships governing the sedimen-
result in a unidirectional seaward flux of sediment. In estuaries, tary processes. An approach of this kind represents a way
because there is unfilled accommodation within the drowned forward (most likely through the techniques of computer mod-
coastal zone, wave and tidal processes produce a landward eling; see section below) that will provide a better ability to
sediment flux from the marine end of the system that supple- predict facies relationships.
ments that from the fluvial end. In addition, the geometry of a
delta tends to favor ebb-tidal dominance while that of an estuary Brackish Ichnology
tends to favor flood-tidal dominance (cf. Friedrichs and Aubrey,
1988). Because estuaries, like other river-mouth coastal environ-
ments, are characterized by brackish-water conditions, the devel-
RECENT AND FUTURE DEVELOPMENT opment of techniques to identify brackish-water deposits using
OF ESTUARINE AND INCISED-VALLEY trace fossils has greatly assisted the recognition of estuarine
FACIES MODELS deposits. In many deposits, distinctive body fossils are either
lacking or poorly preserved, whereas trace fossils are abundant
In this section we first look at the general concept of scientific and preserved in situ. The development of brackish-water ichnol-
models, to identify the current state of evolution of facies models. ogy is a relatively recent field, with early work in the 1980s (e.g.,
We then examine some specific advances in the field of E&IV Wightman et al., 1987) and first-generation summaries published
models and look forward to the approach of the future. in the 1990s (MacEachern and Pemberton, 1994; MacEachern in
Zaitlin et al., 1995). More recent reviews are provided by Pemberton
Development of Scientific Models et al. (2004) and Buatois et al. (2005).
Recent research (e.g., MacEachern and Pemberton, 1994) has
Goodwin (1999) provides an insight into the evolutionary shown a distinctive assemblage of trace fossils for brackish-water
stages in the development of a scientific field such as sedimentol- settings that contrasts strongly with surrounding terrestrial or
ogy (Fig. 44). He identifies an early observation stage that is then fully marine trace-fossil suites (Fig. 45). Additional work by
ESTUARINE AND INCISED-VALLEY FACIES MODELS 219

stressed environments, particularly those subjected to fluctua-


tions in salinity, episodic deposition, variable aggradation rates,
and variability in substrate consistency.”
Recognition of these ichnological characteristics in combina-
tion with the other criteria for distinguishing E&IV systems
given above provides a strong basis for identifying E&IV sys-
tems, even where they exhibit a mud-on-mud or a sand-on-sand
contact with the deposits of other environments. In addition,
careful documentation of ichnofacies assemblages may enable
an internal subdivision of estuarine depositional settings into
bayhead delta, central basin, and barrier components (Fig. 46)
on the basis of a longitudinal gradient of salinity (from nearly
freshwater at the head to nearly marine salinity near the mouth;
cf. MacEachern et al., 1992). A key to the ichnological identifica-
tion of incised-valley deposits is the presence of a firmground
Glossifungites ichnofacies (Fig. 47) that frequently occurs on the
sequence-bounding unconformity at the base of the valley
FIG. 44.—Evolutionary stages in the development of a scientific (MacEachern et al., 1992; Pemberton et al., 1992). It must be
field (after Goodwin, 1999). remembered, however, that all of these ichnological character-
istics may occur in any brackish-water setting and not just
estuaries.
Buatois et al. (1997) has shown that terrestrial trace-fossil assem-
blages in tidal rhythmites can be used to locate the innermost Subdivision of Compound Incised-Valley Fills
tidally influenced freshwater zone of an estuary (see Fig. 10) The
trace-fossil suite of brackish-water environments is characterized A recent advance has been to use detailed compositional data
(MacEachern and Pemberton, 1994) by “a variable and sporadic to subdivide complex, compound valley fills into their constitu-
distribution of burrowing, variability in ichnogenera distribu- ent sequences. Early approaches to incised valleys regarded the
tion, and dominance by simple structures of trophic generalists. fill as an undifferentiated entity, and while later work identified
The suite is dominated by opportunistic suites characteristic of individual components such as bayhead deltas and muddy cen-

FIG. 45.—Comparison of ichnological traces from brackish sediments (left: monospecific Gyrolithes traces) and marine sediments
(right: high-species-diversity traces with Helminthopsis and Chondrites dominant) in the Viking Formation, Western Canada
Sedimentary Basin. (Figure courtesy of James MacEachern.)
220 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 46.—Examples of distinctive ichnofacies from the inner (bay-head delta), middle (central basin), and outer (flood-tidal delta)
regions of an estuary. (After MacEachern and Pemberton, 1994.)

tral basins, the work of Zaitlin et al. (1994) highlighted the return paths from each location in the subsurface, providing
complex nature of many incised-valley deposits as a result of cut continuous coverage and a resulting 3D cube of seismic data
and fill over several sea-level cycles. However, these cycles rather than a 2D slice. Secondly, the 3D method of generating
remain difficult to subdivide, especially in low-accommodation seismic data enables the 3D cube to be sliced horizontally as well
settings such as described in Case Study 4 above, despite being of as vertically. It also allows the 3D cube to be imaged along
prime importance in establishing petroleum reservoir and seal individual reflection horizons, which in turn allows visualization
relationships. Work by Zaitlin et al. (2002) has illustrated how the of complex paleogeomorphological features. A range of seismic
use of a small number of diagnostic petrological components can attributes can be used to highlight aspects of the 3D data (e.g.,
be used to differentiate two cycles and five units of cut and fill in Figs. 24, 33, 49). These include peak-amplitude maps of the
a single formation (Figs. 40–43, 48). Other similar opportunities depositional surface, and classification of the waveforms being
exist to use complementary parameters such as chemostratig- reflected from that surface. These techniques have greatly en-
raphy, heavy minerals, reservoir properties such as pressure and hanced our ability to image E&IV settings (e.g., Zeng et al., 1996;
flow, and remotely sensed electrical properties to identify and Posamentier, 2001; Miall, 2002; Reuter and Watts, 2004) because
subdivide compound valley fills, as well as to determine their the fill of incised valleys frequently differs seismically from the
provenance. Chemostratigraphy, for example, involves the char- surrounding regional sediments. The acoustic-impedance con-
acterization and correlation of strata using major-element and trast at the base of the valley aids further in imaging the valley
trace-element geochemistry and has been used effectively in the container. Finally, the fragmentary coverage of 2D seismic that
North Sea (e.g., Preston et al., 1998) and the Western Canada was ineffective at detecting E&IV facies has been replaced by
Sedimentary Basin (e.g., Ratcliffe et al., 2004). horizontal maps of seismic attributes that are particularly effec-
tive in connecting together the linked reflections that result from
3-D Seismic long, linear coherent features such as channels and valleys (e.g.,
the tributary valleys seen in Figure 49). Increased future use of 3D
Earlier 2D seismic-reflection technology was not effective at seismic processing and enhancement algorithms will be espe-
imaging E&IV systems in the subsurface. This was because the cially powerful for delineating valley networks and longitudinal
frequencies generated by conventional seismic sources were in changes in the nature of the valley-filling deposits.
the range of 20–100 Hz, which is generally not sufficient to
resolve incised valleys with only a few meters to several tens of Numerical Modeling
meters of relief. In addition, 2D seismic collected in single lines
could not provide a regional map of incised-valley distribution, As discussed above regarding scientific models, forward
which typically exhibits a complex regional pattern (e.g., Figs. progress in the field of facies models will require the develop-
26, 35, 41, 49). ment of quantitative techniques to predict the response of E&IV
The advent of 3D seismic changed this scenario in several systems to the dominant processes, and to assess the balance
important ways. Firstly, because 3D-seismic acquisition works between sediment flux and relative sea-level changes. Some
with an array of receivers for each shot location, there are multiple important steps have already been taken in this direction, and
ESTUARINE AND INCISED-VALLEY FACIES MODELS 221

FIG. 47.—Demarcation of incised-valley surfaces by the Glossifungites ichnofacies (from McEachern and Pemberton, 1994). Note
that this ichnofacies is not unique to sequence boundaries.

preliminary results are available from a number of approaches. or echosounder profiles. These earlier acoustic techniques relied on
There is insufficient space to review this field here, but some of the wide-angle single-beam methods with limited spatial coverage.
more interesting approaches are as follows: Results were frequently contoured to give a final representation of
the current marine depositional surface. However, the detailed
(1) The generation of valleys and their fill has been modeled from character of the seafloor remained elusive, and the ability to image
the perspective of landform evolution models (e.g., Willgoose details of the marine depositional surface lagged behind equiva-
et al., 2003; Whipple and Tucker, 2002), fluid mechanics mod- lent land-based approaches such as aerial photography and satel-
els (e.g., Thorne 1994), and alluvial-simulation stratigraphic lite imagery. The development of multibeam sounders, wide-
models (Bridge and Leeder, 1979; Bridge and Mackey, 1993); swath side-scan sonars, and the first seabed returns from 3D
seismic surveys, combined with accurate satellite position fixing,
(2) The question of E&IV preservation has been modeled with a have fundamentally changed our view of the seabed over the past
shoreface-erosion approach (Figs. 50–51) by Cowell et al. twenty years, but particularly over the past five to ten years (e.g.,
(1995), Cowell et al. (1999), and Cowell et al. (2003); Fig. 54). All three of these depth-measuring methods rely on the
propagation of sound waves through the ocean and their reflection
(3) Extensive numerical modelling of estuarine circulation (e.g., from the seabed, providing a marine acoustic image equivalent of
the NOAA model for Chesapeake Bay; NOAA, 2003; Fig. 52) aerial photographs, Landsat images, and digital-elevation models
and sediment transport has been conducted; and for the terrestrial environment. This provides us with our first real
view of what is on the ocean floor at the same degree of resolution
(4) Quantitative relationships have been developed for the bal- as that available on land. Detailed understanding of the modern
ance between river, wave, and tidal power (Fig. 53) and used depositional surface in estuaries enables us to interpret better the
to test the Boyd et al. (1992) coastal classification through the vertical stacking of depositional and erosional surfaces that are
analysis of all major Australian estuaries (Harris et al., 2002). imaged below the seabed in 3D seismic data.
These new views of estuaries have shown us tidal bedforms in
However, these quantitative approaches only address indi- great detail, from the centimeter to the tens-of-meters scale (Fig.
vidual components of the larger system; a full simulation of E&IV 54). They have provided details of separation of flood and ebb
stratigraphy has not yet been attempted. tidal currents, maps of the distribution of the marine flora and
fauna, the nature of deep scour holes, and the release of biogenic
Seabed Imagery and thermogenic gas from pockmarked estuary floors. Deriva-
tion of acoustic backscatter values from side-scan and multibeam
Improved technology for imaging the modern seabed offers data has enabled correlations to be made with sediment grain size
important new insight into marine sedimentary environments. and hence has provided the promise of remotely mapping the
Earlier views of the seabed were derived primarily from individual detailed distribution of sediment texture on the floor of estuarine
soundings, followed more recently by continuous 2D seismic and/ and adjacent shallow-marine areas.
222 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN
ESTUARINE AND INCISED-VALLEY FACIES MODELS 223

FIG. 49.—3D seismic time slice of a Late Pleistocene incised valley from the Java Sea shelf, offshore Indonesia. Note valleys tributary
to the main valley. (From Posamentier, 2001.) Compare with Figure 18.

A LOOK FORWARD—ELEMENTS OF AN E&IV the approach of developing more realistic and quantitative
FACIES MODEL FOR THE TWENTY-FIRST CENTURY facies models outlined above, the following elements represent
important components of an E&IV facies model for the twenty-
The recent advances in E&IV models documented above, first century:
and progress in the field of facies models in general, enable us
to delineate an ideal facies model of the future. Such a model (1) A precise definition of the E&IV system and its morphologi-
would: (1) produce a range of realistic E&IV stratigraphy and cal elements.
facies from a given set of input parameters (see example list
below), (2) identify the preservation potential of the stratigra- (2) A quantitative (digital) database of the geometry and
phy produced in that model, (3) hindcast the input parameters facies of entire systems and their component elements
for a given field example, and (4) predict the rest of the model or from many global examples, both ancient and modern.
example from elements of the component data set. Following This should cover the spectrum of systems and be able to


FIG. 48 (opposite page).—Representative thin sections of the major Basal Quartz units with associated point-count data. Two sets of
ternary diagrams are used to illustrate variations in textural and mineralogical maturity. The upper ternary diagram of each pair
has quartz, chert, and clay-rich grains at the apices and is effective in partitioning the petrographic data into distinctive
populations of mineralogical maturity. The lower ternary diagram of each pair has intergranular, intragranular, and microporosity
pore types at the apices and is used to illustrate porosity fabric and reservoir quality. The representative thin sections are
organized into two cycles (see Fig. 40). Star and triangles represent locations of point-counted samples in the ternary diagrams.
Left photomicrograph in each pair taken in plane light, right photomicrograph of each pair in crossed polars. Magnification 100x.
QTZ = quartz; CH = chert; AR = argillans; P = porosity.
224 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 50.—Transgressive shoreface simulation of Duck, North Carolina, U.S.A., showing retention of a thin estuarine valley fill
(horizontal stripes) after shoreface translation during the last 9 ky (sea-level curve in upper right). From Cowell et al. (1999) and
Cowell et al. (2003).

FIG. 51.—Transgressive shoreface simulation of Haarlem, The Netherlands, showing reworking of shelf deposits into the backbarrier
during transgression. In contrast to the situation shown in Figure 50, almost all of the estuarine sediments (gray color) have been
removed from the shelf but have been preserved behind the aggrading barrier at the present-day shoreline (which marks the
position of maximum transgression and the landward limit of the ravinement surface). From Cowell et al. (1999) and Cowell et
al. (2003).
ESTUARINE AND INCISED-VALLEY FACIES MODELS 225

FIG.52.—Animation of tidal circulation in Chesapeake Bay. (From http://ccmp.chesapeake.org/C3POANIM/). Color bar on right
shows surface tidal current speed in m/s at 0700 on July 10, 2006. Arrows on figure show direction of water transport.

identify “average” or most frequently occurring geom- subset of the processes to describe wave-dominated estuar-
etries and the common internal facies characteristics of ies, for example, or a fluvially eroded valley, while more
each system element. The database should be managed as complex models would be required to describe the response
an open structure able to be accessed by all researchers via of fluvial and estuarine systems to sea-level change or
grid or web-based computing and have a template for incised-valley evolution over a complete sea-level cycle,
common data entry. and to predict the range of subsequent preservation out-
comes. Ideally, computer-modeling software would also
(3) A list of the major processes operating in the E&IV system have an open architecture and be available on line so that
and a description of their dynamic characteristics. Ex- users could simulate parts of the overall system or link
amples of these processes include, but are not limited to, several modules together following the lead of other geo-
plane jet flow in bayhead and tidal deltas, channelized science modeling networks such as www.geoframework.org
flow in inlets and tidal–fluvial channels, wave motion at for internal earth processes.
the seaward margin, and relative sea-level changes through-
out the system. (6) Output models would exhibit a spectrum of 3D examples
spanning the range of natural E&IV variability, together with
(4) A sediment-input component providing sediment volume, a set of “average” models that would describe the most
direction, texture, and composition. These inputs could be frequently occurring combinations of natural parameters
empirical values or derived in turn from models such as (e.g., most common values of wave height, tidal range, valley
climate simulations, wave, and tide predictions. size, rate of sediment supply, and rate of sea-level variation)
and sediment characteristics. Model output would be evalu-
(5) Computer-modeling software developed to simulate the ated on how well it reproduced type field examples.
processes identified in #3 above using inputs of geometry
and facies from #2 and sediment input from #4. For complex The field of facies models in general has had a rapid rise in
systems such as E&IVs, the software models would require knowledge and application over the past forty years, with
a number of linked modules to incorporate the range of estuary and incised-valley models exhibiting a similar rise in
processes present. Early models could utilize a smaller popularity over the past thirteen years. As new strides are
226 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

FIG. 53.—Coastal classification and empirical model testing from Harris et al. (2002). In this approach, the parameters of log [mean
annual fluvial flow] (vertical axis, right) and fluvial discharge (left) have been plotted on a ternary diagram against log [ratio of
tidal power to wave power]. This provides a quantitative test for the Boyd et al. (1992) coastal-classification scheme using all the
river mouths on the Australian coast. Note segregation of major coastal depositional settings.

made to transform the current spectrum of classification mod- ALLEN, J.R.L., 1963, The classification of cross-stratified units, with notes
els into empirical and theoretical models, simulated on com- on their origin: Sedimentology, v. 2, p. 93–114.
puters and tested in the field, further advances to a new level ALLEN, J.R.L., 1964, Studies in fluviatile sedimentation: six cyclothems
of understanding sedimentary depositional systems can be from the Lower Old Red Sandstone, Anglo-Welsh basin: Sedimentol-
anticipated. ogy, v. 3, p. 163–198.
ALLEN, J.R.L., 1965, A review of the origin and characteristics of recent
ACKNOWLEDGMENTS alluvial sediments: Sedimentology, v. 5, p. 89–191.
ALLEN, J.R.L., AND TARLO, L.B., 1963, The Downtonian and Dittonian
The authors wish to acknowledge the many individuals facies of the Welsh Borderland: Geological Magazine, v. 100, p.
and their home institutions who have assisted them over the 129–155.
past twenty years in the development of models for estuary ALLEN, G.P., 1991, Sedimentary processes and facies in the Gironde
and incised-valley depositional systems. In particular we would estuary: a Recent model of macrotidal estuarine systems, in Smith,
like to thank our colleagues for supplying much critical discus- D.G., Reinson, G.E., Zaitlin B.A., and Rahmani, R.A., eds., Clastic
sion and many of the illustrations for this review paper, Tidal Sedimentology: Canadian Society of Petroleum Geologists,
especially Henry Posamentier, James MacEachern, John Suter, Memoir 16, p. 29–40.
Dale Leckie, Peter Roy, Peter Cowell, Peter Harris, Norm ALLEN, G.P., AND POSAMENTIER, H.W., 1993, Sequence stratigraphy and
Corbett, and George Ardies. Funding for our studies has come facies model of an incised valley fill: The Gironde Estuary, France:
from the Natural Sciences and Engineering Research Council Journal of Sedimentary Petrology, v. 63, p. 378–391.
of Canada, the Australian Research Council, ConocoPhillips, ANDERTON, R., 1985, Clastic facies models and facies analysis, in Brenchley,
PanCanadian Energy (now EnCana Corporation), and Exxon- P.J., and Williams, B.J.P., eds., Sedimentology: Recent Developments
Mobil. Ron Boyd would like to acknowledge the support and Applied Aspects: Oxford, U.K., Blackwell, p. 31–47.
provided by the Center for Coastal and Ocean Mapping, Uni- ARDIES, G.W., 1999, Sedimentology, depositional environments and
versity of New Hampshire, while writing this paper on study high resolution sequence stratigraphy of the Horsefly, BAT and
leave. Ellerslie (Basal Quartz) incised valleys, south-central Alberta,
Canada: Unpublished M.Sc. thesis, Queens University, Kingston,
REFERENCES Ontario, 344 p.
ARDIES, G.W., DALRYMPLE, R.W., AND ZAITLIN, B.A., 2001, Examination of the
ALLABY, A., AND ALLABY, M., 1999, Dictionary of Earth Sciences, Second controls on the planform geometry of the BAT (Basal Quartz; Lower
Edition: Oxford, U.K., Oxford University Press. Cretaceous, Lower Mannville) incised-valley system, Western Canada
ESTUARINE AND INCISED-VALLEY FACIES MODELS 227

FIG. 54.—Reson 8101 multibeam data from Portsmouth Harbor, New Hampshire, U.S.A., showing a high-resolution image of
estuarine geomorphology including the channel thalweg (dark blue), an extensive tidal dune field (center) and localized bedrock
outcrops (e.g., right-hand side of channel). Data collected by NOAA as part of the Shallow Survey 2001 Common Data Set (Mayer
and Baldwin, 2001) and processed by the Center for Coastal and Ocean Mapping, University of New Hampshire. 3-D visualization
created using the Fledermaus software suite. Color bar, top right, shows depths in meters below sea level.

Sedimentary Basin (abstract): Seventh International Conference on ASHLEY, G.M., AND SHERIDAN, R.E., 1994, Depositional model for valley fills
Fluvial Sedimentology, University of Nebraska–Lincoln, Program on a passive continental margin, in Dalrymple, R.W., Boyd, R., and
with Abstracts, p. 47. Zaitlin, B.A., eds., Incised-Valley Systems: Origin and Sedimentary
ARDIES, G.W., DALRYMPLE, R.W., and ZAITLIN, B.A., 2002, Controls on the Sequences: SEPM, Special Publication 51, p. 285 p.
geometry of incised valleys in the Basal Quartz unit (Lower creta- BARTHOLDY, J., and PEDERSEN, J.B.T., 2004, Tidalites 2004, Abstracts from the
ceous), Western Canada Sedimentary Basin: Journal of Sedimentary 6th International Conference on Tidal Sedimentology, 2–5 August
Research, v. 72, p. 602–618. 2004, Copenhagen.
ARKELL, W.J., 1933, Jurassic geology of Great Britain: Oxford, U.K., B ELKNAP , D.F., K RAFT , J.C., AND D UNN , R.K., 1994, Transgressive
Clarendon Press, 806 p. valley-fill lithosomes: Delaware and Maine, in Dalrymple, R.W.,
ARNOTT, R.W.C., ZAITLIN, B.A., AND POTOCKI, D., 2000, Geological controls Boyd, R., and Zaitlin, B.A., eds., Incised-Valley Systems: Origin
on reservoir distribution in the Lower Cretaceous Basal Quartz, Chin and Sedimentary Sequences: SEPM, Special Publication 51, p.
Coulee–Horsefly Lake area, south-central Alberta: Bulletin of Cana- 303.
dian Society of Petroleum Geologists, v. 48, p. 212–229. BEST, J.L., AND ASHWORTH, P.J., 1997, Scour in large braided rivers and the
ARNOTT, R.W.C., ZAITLIN, B.A., AND POTOCKI, D., 2002, Stratigraphic re- recognition of sequence stratigraphic boundaries: Nature, v. 387, p.
sponse to sedimentation in an accommodation-limited setting, Lower 275–277.
Cretaceous Basal Quartz, south-central Alberta, Canada: Bulletin of BIGGS, R.B., 1967, The sediments of Chesapeake Bay, in Lauff, G.H., ed.,
Canadian Petroleum Geology, v. 50, p. 92–104. Estuaries: American Association for the Advancement of Science,
ASHLEY, G.M., AND RENWICK, W.H., 1983, Channel morphology and pro- Publication 83, p. 239–260.
cesses at the riverine–estuarine transition, the Raritan River, New BIRD, E.C.F., 2000, Coastal Geomorphology; An Introduction, New York,
Jersey, in Collinson, J.D., and Lewin, J., eds., Modern and Ancient John Wiley & Sons, 322 p.
Fluvial Systems: International Association of Sedimentologists, Spe- BLACKWELDER, E., 1909, The valuation of unconformities: Journal of Geol-
cial Publication 6, p. 207–218. ogy, v. 17, p. 289–300.
228 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

BLOOMER, R.R., 1977, Depositional environments of a reservoir sandstone BUATOIS, L., MANGANO, M.G., MAPLES, C.G., AND LANIER, W.P., 1997, The
in west-central Texas: American Association of Petroleum Geolo- paradox of nonmarine ichnofacies in tidal rhythmites: integrating
gists, Bulletin, v. 61, p. 344–359. sedimentologic and ichnologic data from the late Carboniferous of
BLUM, M.D., 1990, Climatic and eustatic controls on Gulf Coast Plain eastern Kansas, U.S.A.: Palaios, v. 12, p. 467–481.
fluvial sedimentation: an example from the Late Quaternary of the CANT, D.J., AND WALKER, R.G., 1976, Development of a braided-fluvial
Colorado River, Texas, in Armentrout, J.M., and Perkins, B.F., eds., facies model for the Devonian Battery Point Sandstone, Quebec:
Sequence Stratigraphy as an Exploration Tool: Concepts and Prac- Canadian Journal of Earth Sciences, v. 13, p. 102 –119.
tices in the Gulf Coast: SEPM, Gulf Coast Section, 11th Annual CANT, D.J., AND WALKER, R.G., 1978, Fluvial processes and facies sequences
Research Conference, Proceedings, p. 71–83. in the sandy braided South Saskatchewan River, Canada: Sedimen-
BLUM, M.D., 1994, Genesis and architecture of incised valley fill sequences: tology, v. 25, p. 625 –648.
a Late Quaternary example from the Colorado River, Gulf Coastal Plain CLIFTON, H.E., 1982, Estuarine deposits, in Scholle, P.A., and Spearing, D.,
in Texas, in Weimer, P., and Posamentier, H.W., eds., Siliciclastic eds., Sandstone Depositional Environments: American Association
Sequence Stratigraphy: Recent Developments and Applications: Ameri- of Petroleum Geologists, Memoir 31, p. 179–189.
can Association of Petroleum Geologists, Memoir 58, p. 259–283. COLEMAN, J.M., AND PRIOR, D.B., 1980, Deltaic Sand Bodies: American
BLUM, M.D., TOOMEY, R.S., III, AND VALASTRO, S., JR., 1994, Fluvial response Association of Petroleum Geologists, Educational Course Note Series
to Late Quaternary climatic and environmental change, Edwards 15, 171 p.
Plateau, Texas: Paleogeography, Paleoclimatology, Paleoecology, COLEMAN, J.M., AND WRIGHT, L.D., 1975, Modern river deltas: variability of
v. 108, p. 1–21. processes and sand bodies, in Broussard, M.L., ed., Deltas—Models
BLUM, M.D., AND TÖRNQVIST, T.E., 2000, Fluvial responses to climate and for Exploration: Texas, Houston Geological Society, p. 99–149.
sea-level change: a review and a look forward: Sedimentology, v. 47, COOPER, J.A.G., 1988, Sedimentary environments and facies of the sub-
p. 2–48. tropical Mgeni Estuary, southeast Africa: Geological Journal, v. 23, p.
BLUM, M.D., AND VALASTRO, S., JR., 1994, Late Quaternary sedimentation, 59–73.
lower Colorado River, Gulf Coastal Plain of Texas: Geological Society COWELL, P.J., ROY, P.S., CLEVERINGA, J., AND DE BOER, P.L., 1999, Simulating
of America, Bulletin, v. 106, p. 1002–1016. coastal systems tracts using the shoreface translation model: SEPM,
BOSENCE, D.W.J., 1973, Facies relationships in a tidally influenced environ- Special Publication 41, p. 165–175.
ment: a study from the Eocene of the London Basin: Geologie en. COWELL, P.J., ROY, P.S., AND JONES, R.A., 1992, Shoreface translation model—
Mijnbouw, v. 52, p. 63–67. computer-simulation of coastal-sand-body response to sea-level rise:
BOWEN, D.W., AND WEIMER, P., 1997, Reservoir geology of incised valley Mathematical Computer Simulation, v. 33, p. 603–608.
sandstones of the Pennsylvanian Morrow Formation, Southern COWELL, P.J., ROY, P.S., AND JONES, R.A., 1995, Simulation of large-scale
Stateline Trend, Colorado and Kansas, in Shanley, K.W., and Perkins, coastal change using a morphological behavior model: Marine Geol-
B.F., eds., Shallow Marine and Non-Marine Reservoirs—Sequence ogy, v. 126, p. 45–61.
Stratigraphy, Reservoir Architecture and Production Characteris- COWELL, P.J., STIVE, M.J.F., NIEDORODA, A.W., SWIFT, D.J.P., DE VRIEND, D.J.,
tics: SEPM, Gulf Coast Section, Eighteenth Annual Research Confer- BUIJSMAN, M.C., NICHOLLS, R.J., ROY, P.S., KAMINSKY, G.M., CLEVERINGA,
ence, p. 55–66. J., REED, C.W., AND DE BOER, P.L., 2003, The coastal-tract (Part 2):
BOWEN, D.W., AND WEIMER, P., 2003, Regional sequence stratigraphic Applications of aggregated modeling to low-order coastal change,
setting and reservoir geology of Morrow incised-valley sandstones Journal of Coastal Research, v. 19, p. 828–848.
(lower Pennsylvanian), eastern Colorado and western Kansas: Ameri- CRONIN, L.E., 1975, Estuarine Research, Vol. II: Geology and Engineering,
can Association of Petroleum Geologists, Bulletin, v. 87, p. 781–815. New York, Academic Press, Inc., 587 p.
BOYD, R., AND DIESSEL, C.F., 1994, The application of sequence stratigraphy CROSS, T.A., 1988, Controls on coal distribution in transgressive–regres-
to non-marine clastics and coal: Second High Resolution Sequence sive cycles, Upper Cretaceous, Western Interior, U.S.A., in Wilgus,
Stratigraphy Conference, Tremp, Proceedings, p. 13–20. C.K., Hastings, B.S., Posamentier, H.W., Ross, C.A., Van Wagoner, J.,
BOYD, R., DALRYMPLE, R., and ZAITLIN, B.A., 1992, Classification of clastic and Kendall, C.G.St.C., eds., Sea Level Changes: An Integrated Ap-
coastal depositional environments: Sedimentary Geology, v. 80, p. proach: SEPM, Special Publication 42, p. 371–380.
139–150. DAVIS, W.M., 1908, Practical Exercises in Physical Geography: Boston,
BOYD, R., AND HONIG, C., 1992, Estuarine sedimentation on the eastern Ginn and Company, 148 p.
shore of Nova Scotia: Journal of Sedimentary Petrology, v. 62, p. DALRYMPLE, R.W., 1992, Tidal Depositional Systems, in Walker, R.G.,
569–583. and James, N.P., eds., Facies Models: Response to Sea Level Change:
BRIDGE, J.S., AND LEEDER, M.R., 1979, A simulation model of alluvial St. John’s, Newfoundland: Geological Association of Canada, p.
stratigraphy: Sedimentology, v. 26, p. 617–644. 195–218.
BRIDGE, J.S., AND MACKEY, S.D., 1993a, A revised alluvial stratigraphy DALRYMPLE, R.W., 2003, Conference report: SEPM, research conference—
model, in Marzo, M., and Puigdefábregas, C., eds., Alluvial Sedimen- Incised valleys: Images and Processes: SEPM website: www.sepm.org.
tation: International Association of Sedimentologists, Special Publi- DALRYMPLE, R.W., AND ZAITLIN, B.A., 1989, Tidal sedimentation in the
cation 17, p. 319–336. macrotidal, Cobequid Bay–Salmon River estuary, Bay of Fundy:
BROGER, E.J.K., SHYLONYK, G.E., AND ZAITLIN, B.A., 1997, Glauconitic Sand- Canadian Society of Petroleum Geologists, Field Guide, Second Inter-
stone Exploration: A case study from the Lake Newell Project, south- national Symposium on Clastic Tidal Deposits, Calgary, Alberta, 84 p.
ern Alberta, in Pemberton, S.G., and James, D.P., eds., Petroleum DALRYMPLE, R.W., KNIGHT, R.J., ZAITLIN, B.A., AND MIDDLETON, G.V., 1990,
Geology of the Cretaceous Mannville Group, Western Canada: Cana- Dynamics and facies model of a macrotidal sand-bar complex,
dian Society of Petroleum Geologists, Memoir 18, p. 140–168. Cobequid Bay–Salmon River estuary, (Bay of Fundy): Sedimentol-
BROWN, L.F., JR., 1993, Seismic and Sequence Stratigraphy: Its Current ogy, v. 37, p. 577–612.
Status and Growing Role in Exploration and Development (course DALRYMPLE, R.W., MAKINO, Y., AND ZAITLIN, B.A., 1991, Temporal and
notes): New Orleans Geological Society, Short Course No. 5. spatial patterns of rhythmite deposition on mudflats in the macrotidal
BUATOIS, L.A., GINGRAS, M.K., MACEACHERN, J., MANGANO, M.G., ZONNEVELD, Cobequid Bay–Salmon River estuary, Bay of Fundy, Canada, in
J.-P., PEMBERTON, G., NETTO, R.G., AND MARTIN, A., 2005, Colonization Smith, D.G., Reinson, G.E., Zaitlin, B.A., and Rahmani, R.A., eds.,
of brackish-water systems through time: Evidence from the trace- Clastic Tidal Sedimentology: Canadian Society of Petroleum Geolo-
fossil record: Palaios, v. 20, 321–347. gists, Memoir 16, p. 137–160.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 229

DALRYMPLE, R.W., ZAITLIN, B.A., AND BOYD, R., 1992, Estuarine facies mod- carbonate margin during sea level lowstand? Sedimentary Geology,
els: Conceptual basis and stratigraphic implications: Journal of Sedi- v. 157, p. 291–301.
mentary Petrology, v. 62, p. 1130–1146. FISK, H.N., 1944, Geological investigation of the alluvial valley of the
DALRYMPLE, R.W., BOYD, R., AND ZAITLIN, B.A., 1994a, Incised-Valley Sys- lower Mississippi River: Vicksburg, U.S. Army Corps of Engineers,
tems: Origin and Sedimentary Sequences: SEPM, Special Publication Mississippi River Commission, 78 p.
51, 391 p. Fisk, H.N., 1947, Fine-grained alluvial deposits and their effect on Missis-
DALRYMPLE, R.W., BOYD, R., AND ZAITLIN, B.A., 1994b, History of research, sippi River activity: Vicksburg, U.S. Army Corps of Engineers, Mis-
valley types and internal organization of incised-valley systems: sissippi River Commission, 82 p.
introduction to the volume, in Dalrymple, R.W., Boyd, R., and Zaitlin, FLACH, P.D., AND MOSSOP, G., 1985, Depositional environments of the
B.A., eds., Incised-Valley Systems: Origin and Sedimentary Sequences: Lower Cretaceous McMurray Formation, Athabasca Oil Sands,
SEPM, Special Publication 51, p 3–10. Alberta: American Association of Petroleum Geologists, Bulletin, v.
DALRYMPLE, R.W., LECKIE, D.A., AND TILLMAN, R.W., eds., 2006, Incised- 69, p. 1195–1207.
Valleys in Time and Space: SEPM , Special Publication 85, in press. FRIEDRICHS, G.T., AND AUBREY, D.G., 1988, Non-linear tidal distortion in
DAVIES, J.L., 1964, A morphogenic approach to world shorelines: Zeitschrift shallow well mixed estuaries: A Synthesis: Estuarine, Coastal and
für Geomorphology, v. 8, p. 27–42. Shelf Science, v. 27, p. 521–546.
DAVIS, R.A., JR., AND CLIFTON, H.E., 1987, Sea-level change and the preser- GALLOWAY, W.E., 1975, Process framework for describing the morphologic
vation potential of wave-dominated and tide-dominated coastal se- and stratigraphic evolution of deltaic depositional systems, in
quences, in Nummedal, D., Pilkey, O.H., and Howard, J.D., eds., Sea- Broussard, M.L., ed., Deltas—Models for Exploration: Houston Geo-
level Fluctuation and Coastal Evolution: SEPM, Special Publication logical Society, p. 87–98.
41, p. 167–178. GALLOWAY, W.E., AND HOBDAY, D.K, 1996, Terrigenous Clastic Deposi-
DE BOER, P.L., VAN GELDER, A., AND NIO, S.D., 1988, Tide-Influenced Sedi- tional Systems, Second Edition: New York, Springer-Verlag, 489 p.
mentary Environments and Facies: Boston, D. Reidel Publishing GIBLING, M.R., 1991, Sequence analysis of alluvial-dominated cyclothems
Company, 530 p. in the Sydney Basin, Nova Scotia, in Leckie, D.A, Posamentier, H.W.,
DEMAREST, J.M., II, AND KRAFT, J.C., 1987, Stratigraphic record of Quater- and Lovell, B., eds., NUNA Conference on High Resolution Sequence
nary sea levels: implications for more ancient strata, in Nummedal, Stratigraphy: Geological Association of Canada, Calgary, p. 15–19.
D., Pilkey, O.H., and Howard, J.D., eds., Sea-level Fluctuation and GILBERT, G.K. 1880, Land, sculpture, geology of the Henry Mountains
Coastal Evolution: SEPM, Special Publication 41, p. 223–239. (second edition), U.S. Geographical and Geological Survey, Rocky
DOLSON, J., MULLER, D., EVERTS, M.J., AND STEIN, J.A. 1991, Regional paleo- Mountain Region, 160 p.
geographic trends and production, Muddy Sandstone (Lower Creta- GINGRAS, M.K., PEMBERTON, S.G., AND SAUNDERS, T., 1999, The ichnology of
ceous), Central and Northern Rocky Mountains: American Associa- modern and Pleistocene brackish-water deposits at Willapa Bay,
tion of Petroleum Geologists, Bulletin, v. 75, p. 409–435. Washington: variability in estuarine setting: Palaios, v. 14, p. 352–374.
DOLSON, J.C., AND PIOMBINO, J., 1994, Giant proximal foreland basin non- GOODWIN, C.N., 1999, Fluvial classification: Neanderthal necessity or
marine wedge trap: Lower Cretaceous Cutbank Sandstone, Mon- needless normalcy, in Olson, D.S., and Potyondy, J.P., eds., Wild-
tana, in Dolson, J.C., Hendricks, M.L., and Wescott, W.A., eds., land Hydrology: American Water Resources Association, TPS-99-3,
Unconformity-Related Hydrocarbons in Sedimentary Sequences: p. 229–236.
Guidebook for Petroleum Exploration and Exploitation in Clastic and GUILCHER, A., 1967, Origin of sediments in estuaries, in Lauff, G.H., ed.,
Carbonate Sediments: Rocky Mountain Association of Geologists, Estuaries: American Association for the Advancement of Science,
Denver, p. 135–148. Publication 83, p. 149–157.
DONALDSON, A.C., MARTIN, R.H., AND KANES, W.H., 1970, Holocene Guada- GUSTASON, E.R., RYER, R.A., AND ODLAND, S.K. 1986, Unconformities and
lupe Delta of Texas Gulf Coast, in Morgan, J.P., ed., Deltaic Sedimen- facies relationships of the Muddy Sandstone, northern Powder River
tation; Modern and Ancient: SEPM, Special Publication 15, p. 107– Basin, Wyoming and Montana: American Association of Petroleum
137. Geologists, Bulletin, v. 70, p. 1042.
DROSER, M.L., AND BOTTJER, D.J., 1986, Semiquantitative field classification GUSTASON, E.R., WHEELER, D.A., AND RYER, T.A., 1988, Structural control on
of ichnofabric: Journal of Sedimentary Petrology, v. 56, p. 558–559. paleovalley development, Muddy Sandstone, Powder River Basin,
DROSER, M.L., AND BOTTJER, D.J., 1989, Ichnofabric of sandstones deposited Wyoming: American Association of Petroleum Geologists, Bulletin,
in high-energy nearshore environments: measurement and utiliza- v. 72, p. 871.
tion: Palaios, v. 4, p. 598–604. HAMILTON, D., 1979, The high-energy, sand and mud regime of the Severn
EHLERS, T.A., AND CHAN, M.A., 1999, Tidal sedimentology and estuarine estuary, S.W. Britain, in Severn, R.T., Dineley, D., and Hawker, L.E.,
deposition of the proterozoic Big Cottonwood Formation, Utah: eds., Tidal Power and Estuary Management: Albuquerque, Transat-
Journal of Sedimentary Research, v. 69, p. 1169–1180. lantic Arts Incorporated, Colston Paper 30, p. 162–172.
EMERY, K.O., AND STEVENSON, R.E., 1957, Estuaries and lagoons, in Hedgpeth, HAMILTON, W.D., SCOTT, D., MARSDEN, G., AND BAXTER, B., 2001, Structural
J.W., ed., Treatise on Marine Ecology, Vol. 1: Geological Society of and stratigraphic influences on reservoir compartmentalization in
America, Memoir 67, p. 673–750. low accommodation settings: Basal Quartz A Sandstone, Alderson
ETHRIDGE, F.G., AND SCHUMM, S.A., 1978, Reconstructing paleo-channel Lower Mannville “D4D” Pool (abstract): Canadian Society of Petro-
morphologic and flow characteristics: methodology, limitations and leum Geologists, Annual Convention, Calgary, Program and Ab-
assessment, in Miall, A.D., ed., Fluvial Sedimentology: Canadian stracts, p. 150–151.
Society of Petroleum Geologists, Memoir 5, p. 703–721. HAPP, S.C., RITTENHOUSE, G., AND DOBSON, G.C., 1940, Some principles of
ETHRIDGE, F.G., WOOD, L.J., AND SCHUMM, S.A., 1998, Cyclic variables accelerated stream valley sedimentation: U.S. Department of Agricul-
controlling fluvial sequence development: problems and perspec- ture, Technical Bulletin 695.
tives, in Shanley, K.W., and McCabe, P.J., eds., Relative Role of HARMS, J.C., 1966, Stratigraphic traps in valley fill, western Nebraska:
Eustasy, Climate and Tectonics in Continental Rocks: SEPM, Special American Association of Petroleum Geologists, Bulletin, v. 50, p.
Publication 59, p. 17–29. 2119–2149
FIELDING, C.R., TRUEMAN, J.D., DICKENS, G.R., AND PAGE, M., 2003, Anatomy HARRIS, P.T., 1988, Large-scale bedforms as indicators of mutually evasive
of the buried Burdekin River channel across the Great Barrier Reef sand transport and the sequential infilling of wide-mouthed estuar-
shelf: how does a major river operate on a tropical mixed siliciclastic/ ies: Sedimentary Geology, v. 57, p. 273–298.
230 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

HARRIS, P.T., HEAP, A.D., BRYCE, S.M., PORTER-SMITH, R., RYAN, D.A., AND American Association for the Advancement of Science, Publication
HEGGIE, D.T., 2002, Classification of Australian clastic coastal deposi- 83, p. 121–128.
tional environments based on a quantitative analysis of wave, tide JERVEY, M.T., 1988, Quantitative geological modeling of siliciclastic rock
and river power: Journal of Sedimentary Research, v. 72, p. 858–870. sequences and their seismic expression, in Wilgus, C.K., Hastings,
HAYES, B.J.R. 1986, Stratigraphy of the basal Cretaceous Lower Mannville B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van
Formation, southern Alberta and north-central Montana: Bulletin of Wagoner. J.C., eds., Sea Level Research: An Integrated Approach:
Canadian Petroleum Geology, v. 34, p. 30–48. SEPM, Special Publication 42, p. 47–69.
HAYES, B.J.R., CHRISTOPHER, J.E., ROSENTHAL, L., LOS, G., MCKERCHER, B., JOHNSON, H.D., AND LEVELL, B.K., 1995, Sedimentology of a transgressive,
MINKIN, D., TREMBLAY, Y.M., AND FENNEL, J., 1994, Cretaceous Mannville estuarine sand complex: the Lower Cretaceous Woburn Sands (Lower
Group of the Western Canadian Sedimentary Basin, in Mossop, G., Greensand), southern England, in Plint, A.G., ed., Sedimentary Facies
and Shetsen, I., eds., Geological Atlas of the Western Canada Sedi- Analysis—A Tribute to the Research and Teaching of Harold G.
mentary Basin: Canadian Society of Petroleum Geologists, p. 317–334. Reading: International Association of Sedimentologists, Special Pub-
HAYES, M.O., ed., 1969, Coastal Environments—NE Massachusetts and lication 22, p. 17–46.
New Hampshire: University of Massachusetts, Department of Geol- KAY, M., 1951, North American Geosynclines: Geological Society of
ogy, Technical Report #1-CRG, 462 p. America, Memoir 48, 143 p.
HAYES, M.O., 1975, Morphology of sand accumulations in estuaries: an KINDINGER, J.L., BALSON, P.S., AND FLOCKS, J.G., 1994, Stratigraphy of
introduction to the symposium, in Cronin, L.E., ed., Estuarine Re- Mississippi–Alabama Shelf and Mobile River Incised-Valley Sys-
search, Vol. II: New York, Academic Press, p. 3–22. tem, in Dalrymple, R.W., Boyd, R., and Zaitlin, B.A., eds., Incised-
HAYES, M.O, AND SEXTON, W.L., 1989, Modern Clastic Depositional Envi- Valley Systems: Origin and Sedimentary Sequences: SEPM, Special
ronments, South Carolina: American Geophysical Conference, Wash- Publication 51, p. 83–96.
ington, D.C., Field Trip Guidebook T-371, 85 p. KINDLE, E.M., 1917, Recent and fossil ripple-mark: Geological Survey of
HEIN, F.J., AND LANGENBERG, C.W., 2003, Reply to discussion of C.W. Canada Department of Mines, Geological Survey, Museum Bulletin,
Langenberg, F.J. Hein, D. Lawton, and J. Cunningham, Seismic mod- no. 25, p. 1–56.
eling of fluvial–estuarine deposits in the Athabasca oil sands using KLEIN, G. DEVRIES, 1967, Comparison of recent and ancient tidal flat and
ray-tracing techniques, Steepbank River area, northeastern Alberta: estuarine sediments, in Lauff, G.H., ed., Estuaries: American Associa-
Bulletin of Canadian Petroleum Geology, v. 51, p. 354–366. tion for the Advancement of Science, Publication 83, p. 207–218.
HEMPEL, C.G., 1965, Aspects of Scientific Explanation: New York, The Free KROONENBERG, S.B., BADYUKOVA, E.M., STORMS, J.E.A., IGNATOV, E.I., AND
Press, 505 p. KASIMOV, N.S., 2000, A full sea-level cycle in sixty-five years:
HINE, A.C., 1975, Bedform distribution and migration patterns on tidal barrier dynamics of the Caspian Sea: Sedimentary Geology, v. 134,
deltas in the Chatham Harbor Estuary, Cape Cod, Massachusetts, in p. 257–274.
Cronin, L.E., ed., Estuary Research: New York, Academic Press, p. KRYSTINIK, L.F., 1989, Morrow formation facies geometries and reservoir
235–252. quality in compound valley fills, central State Line area, Colorado and
HORNE, J.C., AND FERM, J.C., 1976, Carboniferous depositional environ- Kansas (abstract): American Association of Petroleum Geologists,
ments in the Pocahontas Basin, eastern Kentucky and southern West Bulletin, v. 73, p. 375.
Virginia, Guidebook: University of South Carolina, Department of KRYSTINIK, L.F., AND BLAKENEY, B.A., 1990, Sedimentology of the upper
Geology. Morrow Formation in eastern Colorado and western Kansas, in
HOWARD, A.D., 1967, Drainage analysis in geologic interpretation: a Sonnenberg, S.A., Shannon, L.T., Rader, K., von Drehle, W.F., and
summary: American Association of Petroleum Geologists, Bulletin, v. Gregory, W., eds., Morrow Sandstones of Southeast Colorado and
51, p. 2246–2259. Adjacent Areas: Rocky Mountain Association of Geologists, p. 37–50.
Howard, J.D., Elders, C.A., and Heinbokel, J.F., 1975, Estuaries of the KRYSTINIK, L.F., AND BLAKENEY-DEJARNETT, B.A., 1994, Sedimentology of the
Georgia Coast, U.S.A.: Sedimentology and Biology. II, Regional Upper Morrow Formation in eastern Colorado and western Kansas
animal–sediment characteristics of Georgia estuaries: Sencken- Morrow sandstone (Pennsylvanian) of southeastern Colorado and
bergiana Maritima, v. 7, p. 33–103. Kansas, in Dolson, J.C., Hendricks, M.L., and Wescott, W.A., Uncon-
HOWARD, J.D., AND FREY, R.W., 1973, Characteristics physical and biogenic formity–Related Hydrocarbons in Sedimentary Sequences, Rocky
structures in Georgia estuaries: American Association of Petroleum Mountain Association of Geologists, Reprint, RMAG Sandstone Res-
Geologists, Bulletin, v. 57, p. 1169–1184. ervoirs of the Rocky Mountains, p. 167–180.
HOWARD, J.D., AND FREY, R.W. 1975, Estuaries of the Georgia Coast, U.S.A.: KRYSTINIK, L.F., AND LECKIE, D.A., 2005, Is that channel a valley? Distin-
Sedimentology and Biology, V. Animal–sediment relationships in guishing between distributaries, tidal inlets and paleovalleys: Ameri-
estuarine point bar deposits, Ogeechee River–Ossabaw Sound: can Association of Petroleum Geologists–SEPM Annual Meeting,
Senckenbergiana Maritima, v. 7, p. 181–203. Calgary, Proceedings and Abstracts, p. A76.
HOWARD, R.H., AND WHITAKER, S.T., 1988, Hydrocarbon accumulation in a KULM, L.D., AND BYRNE, J.V., 1967, Sediments of Yaquina Bay, Oregon, in
paleovalley at Mississippian–Pennsylvanian unconformity near Lauff, G.H., ed., Estuaries: American Association for the Advance-
Hardinville, Crawford County, Illinois: a model paleogeomorphic ment of Science, Publication 83, p. 226–238.
trap: Illinois State Geological Survey, Illinois Petroleum 129, 26 p. LAMBIASE, J.J., 1980, Sediment dynamics in the macrotidal Avon River
HOYT, J.H., AND HENRY, V.J., 1965, Significance of inlet sedimentation in estuary, Bay of Fundy: Canadian Journal of Earth Sciences, v. 17, p.
the recognition of ancient barrier islands: Wyoming Geological 1628–1641.
Association, 19th Field Conference Guidebook, p. 190–194. LAND, C.B., JR., 1972, Stratigraphy of the Fox Hills Sandstone and associated
HUBBARD, D.K., 1975, Morphology and hydrodynamics of the Merrimack formations, Rock Springs Uplift and Wamsutter Arch area, Sweetwater
River ebb-tidal delta, in Cronin, L.E., ed., Estuary Research: New County, Wyoming: a shoreline-estuary sandstone model for the Late
York, Academic Press, p. 253–266. Cretaceous: Colorado School of Mines, Quarterly Journal, 67, 69 p.
HUGHES, P., 1958, Tidal mixing in the Narrows of the Mersey Estuary. LANE, E.W., 1935, Stable channels in erodible materials: American Society
Geophysical Journal of the Royal Astronomical Society, v. 1, p. 271– of Civil Engineers, Transactions, v. 63, p. 123–142.
283. Lane, E.W., 1955, The importance of fluvial morphology in hydraulic
Jennings, J.N., and Bird, E.C.F., 1967, Regional Geomorphological Char- engineering: American Society of Civil Engineers, Transactions, v. 81,
acteristics of Some Australian Estuaries, in Lauff, G.H., ed., Estuaries: p. 1–17.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 231

LAUFF, G.H., 1967, ed., Estuaries: American Association for the Advance- MACEACHERN, J.A., AND PEMBERTON, S.G., 1994, Ichnological aspects of
ment of Science, Publication 83, 757 p. incised valley fill systems from the Viking Formation of the West-
LECKIE, D.A., AND SINGH, C., 1991, Estuarine deposits of the Albian Paddy ern Canada Sedimentary Basin, Alberta, Canada, in Dalrymple,
Member (Peace River Formation), and lowermost Shaftsbury Forma- R.W., Boyd, R., and Zaitlin, B.A., eds., Incised-Valley Systems:
tion, Alberta: Journal of Sedimentary Petrology, v. 61, p. 825–850. Origin and Sedimentary Sequences: SEPM, Special Publication 51,
LECKIE, D.A., STANILAND, M.R., AND HAYES, B.J., 1990, Regional maps of the p. 129–158.
Albian Peace River and lower Shaftesbury Formations on the Peace MACEACHERN, J.A., RAYCHAUDHURI, I., AND PEMBERTON, S.G., 1992, Strati-
River Arch, northwestern Alberta and northeastern British Colum- graphic applications of the Glossifungites Ichnofacies: delineating
bia: Bulletin of Canadian Petroleum Geology, v. 38A, p. 176–189. discontinuities in the rock record, I, in Pemberton, S.G., ed., Applica-
LECKIE, D.A., WALLACE-DUDLEY, K.E., VANBESELAERE, N.A., AND JAMES, D.P., tions of Ichnology to Petroleum Exploration: SEPM, Core Workshop
2005, Sedimentation in a low-accommodation setting: nonmarine 17, p. 169–198.
(Cretaceous) Mannville and marine (Jurassic) Ellis Groups, MACNAUGHTON, R.B, DALRYMPLE, R.W., and NARBONNE, G.M., 1997, Early
Manyberries field, southeastern Alberta: American Association of Cambrian braid-delta deposits, MacKenzie Mountains, north-west-
Petroleum Geologists, Bulletin, v. 88, p. 1391–1418. ern Canada: Sedimentology, v. 44: p. 587–609.
LEEDER, M.J., AND GAWTHORPE, R.L., 1987, Sedimentary models for exten- MARTINSEN, R.S., JIAO, Z.S., IVERSON, W.P., AND SURDAM, R.C., 1994, Paleo-
sional tilt-block/half-graben basins, in Coward, M.P., Dewey, J.F., sols and sub-unconformity traps: Examples from the Muddy Sand-
and Hancock, P.L., eds., Continental Extension Tectonics: Geological stone, Powder River basin, Wyoming, in Dolson, J., ed., Unconfor-
Society of London, Special Publication 28, p. 139–152. mity Related Hydrocarbon Exploration and Exploitation in Carbon-
LEIGHTON, V.L., 1997, River metamorphosis in the upper Morrowan Kinsler ate and Clastic Settings: Rocky Mountain Association of Geologists,
field complex, Southwestern Kansas, in Shanley, K.W., and Perkins, p. 119–130.
B.F., eds., Shallow Marine and Non-Marine Reservoirs—Sequence MAYER, L.A., AND BALDWIN, K., 2001, Shallow water survey 2001: Papers
Stratigraphy, Reservoir Architecture and Production Characteristics: based on selected presentations from the second international confer-
SEPM, Gulf Coast Section, Eighteenth Annual Research Conference, ence on high-resolution surveys in shallow water: Marine Technol-
p. 155–169. ogy Society, Journal, v. 35, p. 3–4.
LEOPOLD, L.B., AND WOLMAN, M.G., 1957, River Channel Patterns; Braided, MCCARTHY, P.J., AND PLINT, A.G., 1998, Recognition of interfluve sequence
Meandering, and Straight: U.S. Geological Survey, Professional boundaries: Integrating paleopedology and sequence stratigraphy:
Paper 282-B, 85 p. Geology, v. 26, p. 387–390.
LEOPOLD, L.B., WOLMAN, M.G., and MILLER, J.P., 1964, Fluvial Processes in MCCUBBIN, D.G., 1982, Barrier island and strand plain facies, in Scholle, P.A.,
Geomorphology: San Francisco, Freeman, 522 p. and Spearing, D., eds., Sandstone Depositional Environments: Ameri-
LEROUX, S.M., MACEACHERN, J.A., AND ZAITLIN, B.A., 2001, Ichnological and can Association of Petroleum Geologists, Memoir 31, p. 247–279.
sedimentological analysis of incised valley trends in the Glauconite MELTON, F.A., 1936, An empirical classification of flood-plain streams:
Member and Viking Formation, Central Alberta, Canada (abstract), in Geographical Review, v. 26, p. 593–609.
Mason, J.A., Diffendal, R.F., Jr., and Joeckel, R.M., eds., University of MIALL, A.D., 1977, A review of the braided river depositional environ-
Nebraska–Lincoln, Seventh International Conference on Fluvial Sedi- ment: Earth-Science Reviews, v. 13, p. 1–62
mentology, Program with Abstracts, p. 172 MIALL, A.D., ed., 1978, Fluvial sedimentology: Canadian Society of Petro-
LEVY, J., AND CHRISTIE-BLICK, N., 1994, Neoproterozoic incised valleys of the leum Geologists, Memoir 5, 859 p.
eastern Great Basin, Utah and Idaho, fluvial response to changes in MIALL, A.D., 1985, Architectural-element analysis: a new method of facies
depositional base level, in incised-valley systems, in Dalrymple, R.W., analysis applied to fluvial deposits: Earth-Science Reviews, v. 22, p.
Boyd, R., and Zaitlin, B.A., eds., Incised-Valley Systems: Origin and 261–308.
Sedimentary Sequences: SEPM, Special Publication 51, p. 369–384. MIALL, A.D., 1996, The Geology of Fluvial Deposits: Berlin, Springer, 582 p.
LUKIE, T., 1999, A study of the sedimentology, geochemistry, and MIALL A.D., 2002, Architecture and sequence stratigraphy of Pleistocene
stratigraphic organization of the Lower Cretaceous Horsefly (Basal fluvial systems in the Malay Basin, based on seismic time-slice analy-
Quartz) Valley (Taber–Cutbank), Southern Alberta and Northern sis: American Association of Petroleum Geologists, Bulletin, v. 86, p.
Montana: Unpublished M.Sc. thesis, Queen’s University, Kingston, 1201–1216.
Ontario, 291 p. MYRICK, R.M., AND LEOPOLD, L.B., 1963, Hydraulic geometry of a small tidal
LUKIE, T.D., ARDIES, G.W., DALRYMPLE, R.W., AND ZAITLIN, B.A. 2002, Allu- estuary: U.S. Geological Survey, Professional Paper 422-B, 18 p.
vial architecture of the Horsefly Unit (Basal Quartz) in southern NICHOL, S.L., 1991, Zonation and sedimentology of estuarine facies in an
Alberta and northern Montana: influence of accommodation change incised valley, wave-dominated, microtidal setting, in Smith, D.G.,
and contemporaneous faulting: Bulletin of Canadian Petroleum Ge- Reinson, G.E., Zaitlin, B.A., and Rahmani, R.A., eds., Clastic Tidal
ology, v. 50, p. 73–91. Sedimentology: Canadian Society of Petroleum Geologists, Memoir
LYELL, C., 1830, Principles of Geology, 3 vols.: Murray, London (reprinted 16, p. 41–58.
by Johnson, New York, 1969). NICHOLS, M.M., AND BIGGS, R.B., 1985, Estuaries, in Davis, R.A., Jr., ed.,
MACKIN, J.H., 1937, Erosional history of the Big Horn Basin, Wyoming: Coastal Sedimentary Environments: Second Edition: New York,
Geological Society of America, Bulletin, v. 48, p. 813–894. Springer-Verlag, p. 77–186.
MACKIN, J.H., 1948, Concept of the graded river: Geological Society of NICHOLS, M.M., JOHNSON, G.H., AND PEEBLES, P.C., 1991, Modern sediments
America, Bulletin, v. 59, 463–512 p. and facies model for a microtidal coastal plain estuary, the James
MACEACHERN, J.A., 1999, Ichnology and sedimenology in a sequence estuary, Virginia: Journal of Sedimentary Petrology, v. 61, p. 883–899.
stratigraphic framework: integrated facies models for subsurface NOAA, 2003, http://nauticalcharts.noaa.gov/csdl/op/c3po.html
analysis: Canadian Society of Petroleum Geologists, Short Course OERTEL, G.F., 1975, Ebb-tidal deltas of Georgia estuaries, in Cronin, L.E.,
Notes, 130 p. ed., Estuarine Research, Vol II: New York, Geology and Engineering,
MACEACHERN, J.A., AND PEMBERTON, S.G., 1992, Ichnological aspects of Academic Press, Inc., p. 267–276.
Cretaceous shoreface successions and shoreface variability in the OLSEN, T., STEEL, R., HOGSETH, K., SLAR, T., AND ROE, S.L., 1995, Sequential
Western Interior Seaway of North America, in Pemberton, S.G., ed., architecture in a fluvial succession: sequence stratigraphy in the
Applications of Ichnology to Petroleum Exploration: SEPM, Core Upper Cretaceous Mesaverde Group, Price Canyon, Utah: Journal of
Workshop 17, p. 57–84. Sedimentary Research, v. 65, p. 265–280.
232 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

OOMKENS, E., AND TERWINDT, J.H.J., 1960, Inshore estuarine sediments of PORTER, K.W., AND SONNENBERG, S.A., 1994, Overview of criteria for
the Haringvleit (Netherlands): Geologie en Mijnbouw, v. 39, p. recognition and mapping of valley fills and associated surfaces:
701–710. Lower Cretaceous Muddy Sandstone, Rocky Mountain region, in
PATTISON, S.A., 1991, Crystal, Sundance and Edson valley fill deposits, in Dolson, J.C., Hendricks, M.L., and Wescott, W.A., eds., Unconfor-
Leckie, D.A., Posamentier, H.W., and Lovell, R.R., eds., NUNA Con- mity—Related Hydrocarbons in Sedimentary Sequences: Rocky
ference on High Resolution Sequence Stratigraphy, Banff. Mountain Association of Geolgists, p. 157–166.
PATTISON, S.A.J., AND WALKER, R.G., 1994, Incision and filling of a lowstand POSAMENTIER H.W., 2001, Lowstand alluvial bypass systems: Incised vs.
valley: Late Albian Viking Formation at Crystal, Alberta, Canada: unincised: American Association of Petroleum Geologists, Bulletin,
Journal of Sedimentary Research, v. B64, p. 365–379. v. 85, p. 1771–1793.
PATTISON S.A.J., AND WALKER, R.G., 1998, Multiphase transgressive filling POSAMENTIER, H.W., JERVEY, M.T., and VAIL, P.R., 1988, Eustatic controls on
of an incised valley and shoreface complex, Viking Formation, clastic deposition I—conceptual framework, in Wilgus, C.K., Hastings,
Sundance–Edson area, Alberta: Bulletin of Canadian Petroleum B.S., Kendall, C.G.St.C., Posamentier. H.W., Ross, C.A., and Van
Geology, v. 46, p. 89–105. Wagoner, J.C., eds., Sea Level Changes: An Integrated Approach:
PAYTON, C.E., ed., 1977, Seismic Stratigraphy—Applications to Hydro- SEPM, Special Publication 42, p. 109–124.
carbon Exploration: American Association of Petroleum Geologists, POSAMENTIER, H.W., AND VAIL, P.R., 1988, Eustatic controls on clastic
Memoir 26, 516 p. sedimentation II—Sequence and systems tract models, in Wilgus,
PEIJS-VAN HILTEN, M., GOOD, T.R., AND ZAITLIN, B.A., 1998, Heterogeneity C.K., Hastings, B.S., Ross, C.A., Posamentier, H.W., Van Wagoner, J.,
modeling and geospeudo upscaling applied to waterflood perfor- and Kendall, C.G.St.C., eds., Sea Level Changes: An Integrated Ap-
mance prediction of an incised valley reservoir: Countess YY pool, proach: SEPM, Special Publication 42, p. 125–154.
southern Alberta, Canada: American Association of Petroleum Ge- POSTMA, H., 1967, Sediment transport and sedimentation in the estuarine
ologists, Bulletin, v. 82, p. 2220–2245. environment, in Lauff, G.H., ed., Estuaries: American Association for
PEMBERTON, S.G., SPILA, M., PULHAM, A.J., SAUNDERS, T., MACEACHERN, J.A., the Advancement of Science, Publication 83, p. 158.
ROBBINS, D., AND SINCLAIR, I.K., 2001, Ichnology and Sedimentology of POWELL, J.W., 1875, Exploration of the Colorado River of the West. Wash-
Shallow to Marginal Marine Systems: Geological Association of ington (see also U.S. 43rd Congress, 1st Session, H miscellaneous
Canada, Short Course Notes, no. 15, 343 p. documents 265, 1874).
PEMBERTON, S.G., MACEACHERN, J.A., AND FREY, R.W., 1992, Trace fossil PREDDY, W.S., 1954, The mixing and movement of water in the estuary
facies models: environmental and allostratigraphic significance, in of the Thames: Marine Biology Association, Journal, v. 33, p. 645–
Walker, R.G., and James, N.P., eds., Facies Models: Response to Sea 662.
Level Change: St. John’s, Newfoundland, Geological Association of PRESTON, J., HARTLEY, A., HOLE, M., BUCK, S., BOND, J., MANGE, M., AND STILL,
Canada, 409 p. J., 1998, Integrated whole-rock trace element geochemistry and heavy
PEMBERTON, S.G., AND WIGHTMAN, D.M., 1992, Ichnological characteristics mineral chemistry studies: aids to correlation of continental red-bed
of brackish water deposits, in Pemberton, S.G., ed., Applications of reservoirs in the Beryl Field, U.K. North Sea: Petroleum Geoscience,
Ichnology to Petroleum Exploration: SEPM, Core Workshop 17, p. v. 4, p. 7–16.
141–167. PRITCHARD, D.W., 1967, What is an estuary? Physical viewpoint, in Lauff,
PENCK, A., AND BRÜCKNER, E., 1909, Die Alpen in Eiszeitalter. Leipzig, G.H., ed., Estuaries: American Association for the Advancement of
Tauchnitz. Science, Publication 83, p. 3–5.
PEPPER, J.F., DEWITT, W., JR., AND DEMAREST, D.F., 1954, Geology of the PULHAM, A. J., 1994, The Crusiana field, Llanos basin, eastern Colombia:
Bedford Shale and Berea Sandstone in the Appalachian Basin: U.S. high resolution sequence stratigraphy applied to late Paleocene–
Geological Survey, Professional Paper 259, 109 p. early Oligocene, estuarine, coastal plain and alluvial clastic reser-
PERILLO, G.M.E., ed., 1995, Geomorphology and Sedimentology of Estu- voirs, in Johnson, S., ed., High Resolution Sequence Stratigraphy:
aries: Amsterdam, Elsevier Science, Developments in Sedimentol- Innovations and Applications: University of Liverpool, Liverpool,
ogy, no. 53, 471 p. England, p. 63–68.
PERILLO, G.M.E., PICCOLO, M.C., AND PINO-QUIVIRA, M., 1999, Estuaries of RAHMANI, R.A., 1988, Estuarine tidal channel and nearshore sedimenta-
South America; Their Geomorphology and Dynamics: Berlin, tion of a Late Cretaceous epicontinental sea, Drumheller, Alberta,
Springer, 245 p. Canada, in de Boer, P.L., van Gelder, A., and Nio, S.D., eds., Tide-
PLAYFAIR, J., 1802, Illustrations of the Huttonian theory of the Earth: Influenced Sedimentary Environments and Facies: Boston, D. Reidel
London, Cadel and Davies, (Reprinted by Dover, New York, 1964). Publishing Company, p. 433–481.
PLINK-BJÖRKLUND, P., 2005, Stacked fluvial and tide-dominated estua- RANGER, M.J., AND PEMBERTON, S.G., 1988, Marine influence on the McMurray
rine deposits in high-frequency (fourth-order) sequences of the Formation in the Primrose area, Alberta, in James, D.P., and Leckie,
Eocene Central Basin, Spitsbergen: Sedimentology, v. 52, p. 391– D.A., eds., Sequences, Stratigraphy, Sedimentology; Surface and
428). Subsurface: Canadian Society of Petroleum Geologists, Memoir 15, p.
PLINT, A.G., 2002, Paleovalley systems in the Upper Cretaceous Dunvegan 439–449.
Formation, Alberta and British Columbia: Bulletin of Canadian RATCLIFFE, K.T., WRIGHT, A.M., HALLSWORTH, C., MORTON, A., ZAITLIN, B.A.,
Petroleum Geology, v. 50, p. 277–296. POTOCKI, D., AND WRAY, D.S., 2004, Chemostratigraphic characterization
PLINT, A.G., EYLES, N., EYLES, C.H., AND WALKER, R.G., 1992, Control of sea of fine grained facies in a low accommodation setting: an example from
level change, in Walker, R.G., and James, N.P., eds., Facies Models: the (Lower Cretaceous) Basal Quartz, Southern Alberta: American
Response to Sea Level Change: Geological Association of Canada, St. Association of Petroleum Geologists, Bulletin, v. 88, p. 1419–1432.
John’s, Newfoundland, p. 15–25. REINSON, G.E., CLARK, J.E., AND FOSCOLOS, A.E., 1988, Reservoir geology of
PLINT, A.G., AND WADSWORTH, J.A., 2003, Sedimentology and palaeo- Crystal Viking field, Lower Cretaceous estuarine tidal channel-bay
geomorphology of four large valley systems incising delta plains, complex, south-central Alberta: American Association of Petroleum
Western Canada foreland basin: implications for Mid-Cretaceous Geologists, Bulletin, v. 72, p. 1270–1294.
sea-level changes: Sedimentology, v. 50, p. 1147–1186. REINSON, G.E., 1992, Transgressive barrier island and estuarine systems, in
POR§BSKI, S.J., AND STEEL, R.J., 2003, Shelf-margin deltas: their stratigraphic Walker, R.G., and James, N.P., eds., Facies Models: Response to Sea
significance and relation to deepwater sands: Earth-Science Reviews, Level Change: St. John’s, Newfoundland, Geological Association of
v. 62, p. 283–326. Canada, 409 p.
ESTUARINE AND INCISED-VALLEY FACIES MODELS 233

REINSON, G.E., AND MELOCHE, J.D., 2002, Application of the estuary valley- SHANLEY, K.W., MCCABE, P.J., AND HETTINGER, R.D., 1992, Tidal influences
fill facies model—stratigraphic implications and constraints (ab- in Cretaceous fluvial strata from Utah, USA: a key to sequence
stract): SEPM, Incised Valleys, Images and Processes, Research stratigraphic interpretation: Sedimentology, v. 39, p. 905–930.
Conference, 40 p. SHERWIN, M.D., 1994, Incised channel trends: Glauconitic Member, South-
REUTER, J., AND WATTS D.R., 2004, An ancient river channel system incised ern Alberta, in CSEG-CSPG Joint National Convention 1994, Program
on the Precambrian–Cambrian unconformity beneath Jackson and Extended Abstracts, p. 358.
County, Ohio: American Association of Petroleum Geologists, Bul- SMITH, D.G., 1987, Meandering river point bar lithofacies models: Modern
letin, v. 88, p. 1041–1047. and ancient examples compared, in Ethridge, F.G., Flores, R.M., and
ROCHFORD, D.J., 1951, Studies in Australian estuarine hydrology: Austra- Harvey, M.D., eds., Recent Developments in Fluvial Sedimentology:
lian Journal of Marine and Freshwater Research, v. 2, p. 1–116. SEPM, Special Publication 39, p. 83–91.
ROSENTHAL, L. 1988, Wave-dominated shoreline and incised valley trends: SMITH, D.G., REINSON, G.E., ZAITLIN, B.A., AND RAHMANI, R.A., 1991, Clastic
Lower Cretaceous Glauconitie Formation, west-central Alberta, in Tidal Sedimentology: Canadian Society of Petroleum Geologists,
James, D.J., and Leckie, D.A., eds., Sequences, Stratigraphy Sedi- Memoir 16, 387 p.
mentology; Surface and Subsurface: Canadian Society of Petroleum SNEDDEN, J.W., AND DALRYMPLE, R.W., 1998, Modern shelf sand bodies: an
Geologists, Memoir 15, p. 207–220. integrated hydrodynamic and evolutionary model, in Bergman, K.,
ROSS, G.M., EATON, D.W., BOERNER, D.E., AND CLOWES, R.M., 1997, Geolo- and Snedden, J.W., eds., Isolated Shallow Marine Sandbodies: Se-
gists probe buried craton in Western Canada: EOS, Transactions, quence Stratigraphic Analysis and Sedimentologic Interpretation:
American Geophysical Union, v. 78, p. 493–494. SEPM, Special Publication 64, p. 13–28.
ROY, P.S., 1977, Does the Hunter River supply sand to the New South SOLLAS, W.J., 1883, The estuaries of the Severn and its tributaries: Geologi-
Wales coast today? Royal Society of New South Wales, Journal and cal Society of London, Quarterly Journal, v. 39, p. 611–626.
Proceedings, v. 110, p. 17–24. STEERS, J.A., 1967, Geomorphology and coastal processes, estuaries: Ameri-
ROY, P.S., 1984, New South Wales estuaries: their origin and evolution, can Association for the Advancement of Science, Washington, D.C.,
in Thom, B.G., ed., Coastal Geomorphology in Australia: New York, Publication 83, p. 100.
Academic Press, p. 99–121. SUMMERFIELD, M.A., 1985, Plate tectonics and landscape development on
ROY, P.S., THOM, B.G., AND WRIGHT, L.D., 1980, Holocene sequences on an the African continent, in Morisawa, M., and Hack, J., eds., Tectonic
embayed high energy coast: an evolutionary model: Sedimentary Geomorphology: Boston, Allen and Unwin, p. 27–51.
Geology, v. 26, p. 1–19. SUTER, J.R., AND BERRYHILL, H.L., 1985, Late Quaternary shelf margin deltas,
RUSNAK, G.A., 1967, Rates of sediment accumulation in modern estuar- Northwest Gulf of Mexico: American Association of Petroleum Ge-
ies, in Lauff, G.H., ed., Estuaries: American Association for the ologists, Bulletin, v. 69, p. 77–91.
Advancement of Science, Publication 83, p. 180–184. SUTER, J.R., BERRYHILL, H.L., and PENLAND, S., 1987, Late Quaternary sea-
RUSSELL, R.J., 1967, Origins of estuaries, in Lauff, G.H., ed., Estuaries: level flucturations and depositional sequences, southwest Louisiana
American Association for the Advancement of Science, Publication continental shelf, in Nummedal, D., Pilkey, O.H., and Howard, J.D.,
83, p. 93–99. eds., Sea-Level Fluctuation and Coastal Evolution: SEPM, Special
RUST, B.R., 1978a, A classification of alluvial channel systems, in Miall, Publication 41, p. 199–219.
A.D., ed., Fluvial Sedimentology: Canadian Society of Petroleum SUTER, J.R., AND CLIFTON, H.E., 1999, The Shannon Sandstone and isolated
Geologists, Memoir 5, p. 187–198. linear sand bodies: interpretations and realizations, in Bergman, K.,
RUST, B.R., 1978b, Depositional models for braided alluvium, in Miall, and Snedden, J.W., eds., Isolated Shallow Marine Sandbodies: Se-
A.D., ed., Fluvial Sedimentology: Canadian Society of Petroleum quence Stratigraphic Analysis and Sedimentologic Interpretation:
Geologists, Memoir 5, p. 605–625. SEPM, Special Publication 64, p. 321–356.
SCHUBEL, J.R., AND HIRSCHBERG, D.J., 1978, Estuarine graveyard and cli- SYVITSKI J.P.M., AND DAUGHNEY, S., 1992, DELTA2—Delta-progradation
matic change, in Wiley, M., ed., Estuarine Processes: New York, and basin filling: Computers & Geoscience, v. 18, p. 839–897.
Academic Press, p. 285–303. TERWINDT, J.H.J., 1963, Excursion Rhine–Meuse Estuary (Haringvliet): Sixth
SCHUCHERT, C., 1927, Unconformities as seen in disconformities and International Sedimentological Congress, Amsterdam, Guidebook.
diastems: American Journal of Science, v. 13, p. 260–262. THOMAS, M.A., AND ANDERSON, J.B., 1994, Sea-level controls on the facies
SCHUMM, S.A., 1972, River Morphology: Stroudsburg, Pennsylvania, architecture of the Trinity/Sabine incised-valley system, Texas conti-
Dowden, Hutchinson & Ross, Benchmark Papers in Geology, 429 p. nental shelf, in Dalrymple, R.W., Boyd, R., and Zaitlin, B.A., Incised-
SCHUMM, S.A., 1977, The Fluvial System: New York, Wiley, 338 p. Valley Systems: Origin and Sedimentary Sequences: SEPM, Special
SCHUMM, S.A., 1993, River response to baselevel change: implications for Publication 51, p. 63–82.
sequence stratigraphy: Journal of Geology, v. 101, p. 279–294. THOMAS, R.G., SMITH, D.G., WOOD, J.M., VISSER, J., CALVERLEY-RANGE, E.A.,
SCHUMM, S.A, AND ETHRIDGE, F.G., 1994, Origin, evolution and morphol- AND KOSTER, E.H., 1987, Inclined heterolithic stratification—termi-
ogy of fluvial valleys, in Dalrymple, R.W., Boyd, R., and Zaitlin, nology, description, interpretation and significance: Sedimentary
B.A., eds., Incised-Valley Systems: Origin and Sedimentary Se- Geology, v. 53, p. 123–179.
quences: SEPM, Special Publication 51, p. 11–28. THORNE, J., 1994, Constraints on riverine valley incision and the response
SCHUMM, S.A., AND KHAN, H.R., 1972, Experimental study of channel to sea-level change based on fluid mechanics, in Dalrymple, R.W.,
patterns: Geological Society of America, Bulletin, v. 83, p. 1755– Boyd, R., and Zaitlin, B.A., Incised-Valley Systems: Origin and
1770. Sedimentary Sequences: SEPM, Special Publication 51, p. 29–44.
SEIVER, R. 1951, The Mississippian–Pennsylvanian unconformity in south- TÖRNQVIST, T.E., 1993, Holocene alternation of meandering and anasto-
ern Illinois: American Association of Petroleum Geologists, Bulle- mosing fluvial systems in the Rhine-Mesue delta (central Nether-
tin, v. 35, p. 542–581. lands) controlled by sea-level rise and subsoil erodibility: Journal of
SHANLEY, K.W., AND MCCABE, P.J., 1991, Predicting facies architecture Sedimentary Petrology, v. 63, p. 683–693.
through sequence stratigraphy: Journal of Geology, v. 101, p. 279– TWENHOFEL, W.H., 1932, Treatise on Sedimentation, Second Edition,
294. 1961, New York, Dover Publications, 926 p.
SHANLEY, K.W., AND MCCABE, P.J., 1994, Perspectives on the sequence VALLIANOS, L., 1975, A recent history of Masonboro Inlet, North Carolina,
stratigraphy of continental strata: American Association of Petro- in Cronin, L.E., ed., Estuarine Research, vol. II. Geology and Engi-
leum Geologists, Bulletin, v. 78, p. 544–568. neering: New York, Academic Press, Inc.
234 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN

VAN STRAATEN, L.M.J.U., 1952, Biogene textures and the formation of WELLNER, R.W., AND BARTEK, L.R., 2003, The effect of sea level, climate, and
shell beds in the Dutch Wadden Sea: Koninklijke Nederlandse shelf physiography on the development of incised-valley complexes:
Akademie van Wetenschappen, Proceedings, Series B., 55, p. 500– A modern example from the East China Sea: Journal of Sedimentary
516. Research, v. 73, p. 926–940.
VAN STRAATEN, L.M.J.U., 1954a, Sedimentology of Recent tidal flat de- WESCOTT, W.A., 1993, Geomorphic thresholds and complex response of
posits and the Psammites du Condroz (Devonian): Geologie en fluvial systems—some implications for sequence stratigraphy: Ameri-
Mijnbouw, v. 16, p. 25–47. can Association of Petroleum Geologists, Bulletin, v. 77, p. 1208–1218.
Van Straaten, L.M.J.U., 1954b, Composition and structure of Recent WHIPPLE, K.X., AND TUCKER, G.E., 2002, Implications of sediment flux
marine sediments in the Netherlands: Leidse Geologie Medelingen, dependent river incision models for landscape evolution: Journal of
v. 19, p. 1–110. Geophysical Research, v. 107, no. B2, 10.1029/2000JB000044, ETG 3.1–
VAN STRAATEN, L.M.J.U., 1961, Sedimentation in tidal flat areas: Alberta 3.20.
Society of Petroleum Geologists Journal, 9, p. 203–226. WIGHTMAN, D.M., PEMBERTON, S.G., AND SINGH, C., 1987, Depositional
VAN WAGONER, J.C., MITCHUM, R.M., CAMPION, K.M., AND RAHMANIAN, modeling of the Upper Mannville (Lower Cretaceous), east-central
V.D., 1990, Siliciclastic Sequence Stratigraphy in Well Logs, Cores, Alberta, implications for the recognition of brackish water deposits:
and Outcrops: Concepts for High-Resolution Correlation of Time in Tillman, R.W., and Weber, K.J., eds., Reservoir Sedimentology:
and Facies: American Association of Petroleum Geologists, Meth- SEPM, Special Publication 40, p. 189–220.
ods in Exploration Series, no. 7, 55 p. WILGUS, C.K., HASTINGS, B.S., POSAMENTIER, H.W., ROSS, C.A., VAN WAGONER,
VAN WAGONER, J.C., NUMMEDAL, D., JONES, C.R., TAYLOR, D.R., JENNETTE, J., AND KENDALL, C.G.ST.C., eds., 1988, Sea Level Changes: An Inte-
D.C., and RILEY, G.W. 1991, Sequence Stratigraphy Applications to grated Approach: SEPM, Special Publication 42, 407 p.
Shelf Sandstone Reservoirs: American Association of Petroleum WILLGOOSE, G.R., HANCOCK, G.R., AND KUCZERA, G.A., 2003, A framework
Geologists, Field Conference Guidebook, Tulsa. for the quantitative testing of landform evolution models, in Wilcock,
VAN WAGONER, J.C., POSAMENTIER, H.W., MITCHUM, R.M., VAIL, P.R., SARG, P.R., and Iverson, R.M., eds., Predictions in Geomorphology: Ameri-
J.F., LOUTIT, T.S., AND HARDENBOL, J., 1988, An overview of the funda- can Geophysical Union, Washington, D.C., p. 195–216.
mentals of sequence stratigraphy and key definitions, in Wilgus, WOOD, J.M., AND HOPKINS, J.C., 1989, Reservoir sandstone bodies in estua-
C.K., Hastings, B.S., Posamentier, H.W., Ross, C.A., Van Wagoner, rine valley fill: Lower Cretaceous Glauconitic Member, Little Bow
J., and Kendall, C.G.St.C., eds., Sea Level Changes: An Integrated Field, Alberta, Canada: American Association of Petroleum Geolo-
Approach: SEPM, Special Publication 42, p. 39–45. gists, Bulletin, v. 73, p. 1361–1382.
WADSWORTH, J., BOYD, R., DIESSEL, C., LECKIE, D., AND ZAITLIN, B.A., 2002, WOOD, J.M., AND HOPKINS, J.C., 1992, Traps associated with paleo-valleys
Stratigraphic style of coal and non-marine strata in a tectonically and interfluves in an unconformity bounded sequence: Lower Creta-
influenced intermediate accommodation setting: the Mannville ceous Glauconitic Memer, southern Alberta, Canada: American As-
Group of the Western Canadian Sedimentary Basin, south-central sociation of Petroleum Geologists, Bulletin, v. 76, p. 904–926.
Alberta: Bulletin of Canadian Petroleum Geology, v. 50, p. 507– Wood, L.J., Ethridge, F.G., and Schumm, S.A., 1993, An experimental
541. study of the influence of subaqueous shelf angles on coastal plain
WALKER, R.G., ed., 1979, Facies Models: Geological Association of Canada, and shelf deposits, in Weimer, P., and Posamentier, H.W., eds.,
St. John’s, Newfoundland, Geoscience Canada Reprint Series 1. Siliciclastic Sequence Stratigraphy: American Association of Petro-
WALKER, R.G., ed., 1984a, Facies Models, Second Edition: Geological leum Geologists, Memoir 58, p. 381–391.
Association of Canada, St. John’s, Newfoundland, Geoscience WOODROFFE, C.D., CHAPPELL, J.M.A., THOM, B.G., AND WALLENSKY, E., 1989,
Canada Reprint Series 1. Depositional model of a macrotidal estuary and flood plain, South
WALKER, R.G., ed., 1984b, General introduction: facies, facies sequences and Alligator River, Northern Australia: Sedimentology, v. 36, p. 737–756.
facies models, in Walker, R.G., ed., Facies Models, Second Edition: WOOLFE, K.J., LARCOMBE, P., NAISH, T., and PURDON, R.G., 1998, Lowstand
Geological Association of Canada, Toronto, Geoscience Canada Re- rivers need not incise the shelf: an example from the Great Barrier
print Series 1, p. 1–9. Reef, Australia, with implications for sequence stratigraphic models:
WALKER, R.G., 1992, Facies, facies models and modern stratigraphic Geology, v. 26, p. 75–78.
concepts, in Walker, R.G., and James, N.P., eds., Facies Models: WRIGHT, L.D., 1985, River deltas, in Davis, R.A., Jr., ed., Coastal Sedimen-
Response to Sea Level Change: St. John’s, Newfoundland, Geologi- tary Environments, Second Edition: New York, Springer-Verlag, p.
cal Association of Canada, p. 1–14. 1–76.
WALKER, R.G., AND JAMES, N.P., 1992 eds., Facies Models: Response to Sea WRIGHT, L.D., COLEMAN, J.M., AND THOM, B.G., 1973, Processes of chan-
Level Change: St. John’s, Newfoundland, Geological Association of nel development in a high-tide-range environment: Cambridge
Canada, 409 p. Gulf–Ord River delta, Western Australia: Journal of Geology, v. 81,
WANLESS, H.R., AND WELLER, J.M., 1932, Correlation and extent of Pennsyl- p. 15–41.
vanian cyclothems. Geological Society of America, Bulletin, v. 43, p. WRIGHT, V.P., AND MARRIOTT, S.B., 1993, The sequence stratigraphy of
1003–1016. fluvial depositional systems: the role of floodplain sediment storage:
WASHINGTON, P.A., AND CHISICK, S.A., 1994, Estuarine facies models: con- Sedimentary Geology, v. 86, p. 203–210.
ceptual basis and stratigraphic implications—Discussion: Journal of ZAITLIN, B.A., 2003. Recent advances in hydrocarbon exploration and
Sedimentary Research, v. B64, p. 74–75. exploitation techniques for incised valley systems: examples from the
WAY, J.N., O’MALLEY, P.J., SUTTNER, L.J., and FURER, L.C., 1998, Tectonic Lower Cretaceous of the Western Canada Sedimentary Basin (ab-
controls on alluvial systems in a distal foreland basin: the Lakota and stract): American Association of Petroleum Geologists–SEPM Annual
Cloverly Formations (Early Cretaceous) in Wyoming, Montana and Convention, Salt Lake City, Utah, Proceedings and Abstracts, p. A187.
South Dakota, in Shanley, K., and McCabe, P., eds., Relative Role of ZAITLIN, B.A., DALRYMPLE, R.W., AND BOYD, R., 1994, The stratigraphic
Eustasy, Climate, and Tectonism in Continental Rocks: SEPM, Special organization of incised-valley systems associated with relative sea-
Publication 59, 145 p. level change, in Dalrymple, R.W., Boyd, R., and Zaitlin, B.A., eds.,
WEIMER, R.J., 1984, Relation of unconformities, tectonics, and sea-level Incised-Valley Systems: Origin and Sedimentary Sequences: SEPM,
changes, Cretaceous of the Western Interior, U.S.A., in Schlee, J.S., ed., Special Publication 51, p. 45–60.
Interregional Unconformities and Hydrocarbon Accumulation: ZAITLIN, B.A, DALRYMPLE, R.W., BOYD, R., LECKIE, D., AND MACEACHERN, J.,
American Association of Petroleum Geologists, Memoir 36, p. 7–35. 1995, The Stratigraphic Organisation of Incised Valley Systems, Im-
ESTUARINE AND INCISED-VALLEY FACIES MODELS 235

plications to Hydrocarbon Exploration and Production with Ex-


amples from the Western Canada Sedimentary Basin: Canadian
Society of Petroleum Geologists, Calgary, Alberta, 189 p.
ZAITLIN, B.A., GRIFFITH, L., HEUBSCH, H., LEGGITT, S., DUFRESNE, D., POTOCKI,
D., COX, W., SQUIRES, J., AND SMITH, I., 1998, Lathom “A” Pool: an
example of a Lower Cretaceous compound incised valley reservoir
from the western interior (abstract): American Association of Petro-
leum Geologists, Annual Convention, Program and Abstracts.
ZAITLIN, B.A., AND SHULTZ, B.C., 1984, An estuarine-embayment fill model
from the Lower Cretaceous Mannville Group, west-central
Saskatchewan, in Scott, D.F., and Glass, D.J., eds., Mesozoic of Middle
North America: Canadian Society of Petroleum Geologists, Memoir
9, p. 455–469.
ZAITLIN, B.A., AND SHULTZ, B.C., 1990, Wave-influenced estuarine sand
body, Senlac heavy oil pool, Saskatchewan, Canada, in Barwis, J.H.,
McPherson, J.G., and Studlick, J.R.J., eds., Sandstone Petroleum Res-
ervoirs: New York, Springer-Verlag, p. 363–387.
ZAITLIN, B.A., WARREN, M.J., POTOCKI, D., ROSENTHAL, L., AND BOYD, R., 2002,
Depositional styles in a low accommodation foreland basin setting:
an example from the Basal Quartz (Lower Cretaceous), southern
Alberta: Bulletin of Canadian Petroleum Geology, v. 50, p. 31–72.
ZENG, H.L., BACKUS, M.M., BARROW, K.T., AND TYLER, N., 1996, Facies
mapping from three-dimensional seismic data: potential and guide-
lines from a Tertiary sandstone–shale sequence model, Powderhorn
Field, Calhoun County, Texas: American Association of Petroleum
Geologists, Bulletin, v, 80, p. 16–46.
236 RON BOYD, ROBERT W. DALRYMPLE AND BRIAN A. ZAITLIN
DELTAS 237

DELTAS
JANOK P. BHATTACHARYA
Robert E. Sheriff Professor of Sequence Stratigraphy, Geosciences Department, SR1 Rm. 312,
University of Houston, 4800 Calhoun Rd., Houston, Texas 77204-5007, U.S.A.
e-mail: jpbhattacharya@uh.edu

ABSTRACT: Deltas are discrete shoreline protuberances formed where a river enters a standing body of water and supplies sediments more
rapidly than they can be redistributed by basinal processes, such as tides and waves. In that sense, all deltas are river-dominated and deltas
are fundamentally regressive in nature. The morphology and facies architecture of a delta is controlled by the proportion of wave, tide, and
river processes; the salinity contrast between inflowing water and the standing body of water, the sediment discharge and sediment caliber,
and the water depth into which the river flows. The geometry of the receiving basin (and proximity to a shelf edge) may also have an
influence. The simple classification into river-, wave-, and tide-dominated end members must be used with caution because the number
of parameters that control deltas is more numerous.
Other depositional environments, such as wave-formed shorefaces or barrier-lagoons can form significant components of larger wave-
influenced deltas, but conversely smaller bayhead or lagoonal deltas can form within larger barrier-island or estuarine systems. As deltas
are abandoned and transgressed they may also be transformed into another depositional systems (e.g., transgressive barrier–lagoon system
or estuary). Delta plains also contain distributary river channels and their associated floodplains and bays, which can equally be classified
as both fluvial and deltaic environments.
Sharp-based blocky sandstones, tens of meters up to about a hundred meters thick, within many ancient mid-continent deltas have
routinely been interpreted in the rock record as distributary channels, although many of these examples are now reinterpreted as incised
fluvial valleys. Distributary channels may show several orders of sizes and shapes as they bifurcate downstream around distributary-
mouth bars. Bifurcation is inhibited in strongly wave-influenced deltas, resulting in relatively few terminal distributary channels and
mouth bars flanked by extensive wave-formed sandy barriers or strandplain deposits. In shallow-water river-dominated deltas, tens to
hundreds of shallow, narrow and ephemeral terminal distributary channels can form intimately associated with mouth bars that form
larger depositional lobes. Tides appear to stabilize distributary channels for hundred to thousands of years, inhibiting avulsion and delta
switching.
As deltas prograde they form upward-coarsening facies successions, as sandy mouth bars and delta-front sediments build over muddy
deeper-water prodelta facies. Deltas display a distinct down-dip clinoform cross-sectional architecture. Many large muddy deltas show
separate clinoforms, the first at the active sandy delta front and the second on the muddy shelf. Along-strike facies relationships may be
less predictable and depositional surfaces may dip in different directions. Overlapping delta lobes typically result in lens-shaped
stratigraphic units that exhibit a mounded appearance.
All modern deltas grade updip from marine into non marine environments, and Walther’s Law predicts that deltas should show a marine
to nonmarine transition as they prograde. However, in many low-accommodation settings, topset alluvial or delta-plain facies can be
removed or reworked by wave or tidal erosion during transgression, resulting in top-eroded deltas. Historically, some of these top-eroded
deltas have been interpreted as distal shelf deposits, not related to shoreline processes. Sequence stratigraphic concepts, however, allow
facies observations to be placed within a larger context of controlling allocyclic mechanisms which allow the correct interpretation of larger
delta systems of which only small remnants may be preserved.

WHAT ARE DELTAS, plain. Better prediction of deltas, the critical boundary between
AND WHY ARE THEY IMPORTANT? land and sea, is needed.
From the economic perspective, deltas have been estimated to
Much of the sediment transferred from land to sea is carried host close to 30% of all of the world’s oil, coal, and gas deposits
by rivers and deposited at the shoreline in the form of deltas. (Tyler and Finley, 1991). Much of these resources are in areas that
About 25% of the world’s population live on deltaic coastlines have been productive for many years, such as the oil and gas
and wetlands (Syvitski et al., 2005). Prediction of growth and deposits of the Cretaceous Interior Seaway and the Gulf of
decay of modern deltas is critical in areas such as Louisiana and Mexico and the Carboniferous coal deposits of the UK and the
in much of Asia where rampant dam building has caused an Eastern USA. However, as production declines and global energy
immense decrease in discharge of freshwater to the world’s needs continue to grow, new and better facies models will be
oceans, resulting in enormous stresses to these coastal ecosys- required to improve the extraction of oil and gas. Significant
tems as they experience subsidence and land loss (Vörösmarty et fresh-water resources also occur in delta deposits, and exploita-
al., 1997). Sixty percent of the world’s rivers are affected by tion of these aquifers requires robust facies models for deltas.
reservoirs (Syvitski et al., 2005). The USA National Inventory of This paper reviews deltaic facies models from the sedimento-
Dams shows 38,100 dams over 6 m high blocking the flow in the logical to the regional stratigraphic perspective. It begins by
Mississippi River drainage basin alone. The muddy deltaic coast- presenting a brief historical overview of delta studies and then
line immediately south of Bangkok, Thailand, for example, is proceeds with defining what a delta is, describing the various
retreating at 12 m per year as a consequence of the extreme components of deltaic depositional systems, in terms of typical
demands made by a huge population on the precious freshwater sub-environments and facies successions, and then finishes with
resource supplied by the Chao Phraya River (Vongvisessomjai, the larger perspective provided by facies-architecture and se-
1990). Similar problems exist along the entire Indo-Gangetic delta quence-stratigraphic studies.

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 237–292.
238 JANOK P. BHATTACHARYA

Historical Background
The concept of a delta dates back to Herodotus (c. 400 BC),
who recognized that the alluvial plain at the mouth of the Nile
had the form of the capital Greek letter ∆ (Fig. 1). The first study
of ancient deltas was that of Gilbert (1885), who described Pleis-
tocene fresh-water gravelly deltas in Lake Bonneville, Utah.
Gilbert recognized a basic threefold subdivision of delta deposits
into topset, foreset, and bottomset units (Fig. 2), a terminology
that remains in use to this day. Barrell (1912) extended these
subdivisions to the much larger scale of the Devonian Catskill
wedge in the Appalachians, and provided the first explicit defi-
nition of the essential features of a delta as

“… a deposit partly subaerial built by a river into or against a


body of permanent water. The outer and lower parts are necessar-
ily constructed below water level, but its upper and inner surface
must be land maintained or reclaimed by the river building from
the sea. A delta, therefore, consists of a combination of terrestrial
and marine, or at least lacustrine strata, and differs from other
modes of sedimentation in this respect” (Barrell, 1912, p. 381).

Barrell considered the recognition of associated nonmarine


facies crucial in distinguishing ancient deltas from estuaries. This
criterion is no longer required because in deltas deposited during
times of falling sea level (e.g., lowstand, forced regressive, or
falling-stage systems tracts), subaerial topset facies may either
not be deposited or may be eroded during subsequent transgres-
sion, yielding “top-truncated” delta deposits (Plint, 1988;
Posamentier et al., 1992; Hart and Long, 1996; Bhattacharya and FIG. 2.—Cross-sectional facies architecture and vertical facies
Willis, 2001; Martinsen, 2003). succession of a delta showing threefold subdivision into
Our understanding of modern deltas developed rapidly dur- topset, foreset, and bottomset strata. From Elliott (1986), after
ing the last fifty years, beginning with work on the Mississippi Gilbert (1885) and Barrell (1912).
Delta published in the 1950s and early 1960s (e.g., Shepard et al.,
1960). Scruton (1960) recognized that deltas are essentially cyclic
in nature and consist of a progradational, “constructive phase” usually followed by a thinner retrogradational “destructive phase”
coinciding with delta abandonment. He also illustrated a vertical
“deltaic sequence” (Scruton, 1960) of coarsening- and sandier-
upward facies related to progradation of bottomset, foreset, and
topset strata (Fig. 3).
The Gulf Coast region of the U.S.A. (Florida to Texas) histori-
cally has been an important focus for research on modern and
ancient deltas, primarily because of the economic importance of
deltas as oil and gas reservoirs. Coleman and Wright (1975)
compiled a global data base of 34 modern deltas and developed
a six-fold classification based on sand distribution patterns (Fig.
4) with accompanying “typical” vertical facies profiles. One of the
most widely used classification schemes is that of Galloway
(1975), who subdivided deltas according to the dominant pro-
cesses controlling their morphology: rivers, waves, and tides
(Fig. 5). These two studies emphasized the importance of the
overall shape of a sediment body in defining the type of delta,
although this has recently come under some criticism (Dominguez,
1996; Rodriguez et al., 2000; Fielding et al., 2005a). A spate of
research, focused on coarser-grained, high-latitude delta sys-
tems, led to an appreciation of the importance of grain size, water
depth, and feeder type as controlling variables on delta type
FIG. 1.—Environments and facies in the modern Nile delta. Only (Colella and Prior, 1990; Postma, 1990; Orton and Reading, 1993).
the Rosetta and Damietta Branches are presently active. Stipple Improvements in seismic and side-scan sonar imaging led to
indicates older reworked delta sands (Scheihing and Gaynor, the recognition of regional-scale synsedimentary deformation in
1991) rather than active “sand plumes” (Coleman et al., 1981). the subaqueous parts of modern deltas (Coleman et al., 1983;
Since construction of the Aswan Dam, water and sediment Winker and Edwards, 1983). Similar features have now been
discharge to the delta have decreased, and the entire delta is recognized in outcrops of several ancient deltas (Nemec et al.,
undergoing transgression. From Bhattacharya and Walker, 1988; Martinsen, 1989; Pulham, 1989; Bhattacharya and Davies,
(1992) based on Fisher et al. (1969) and Sestini (1989). 2001, 2004; Wignall and Best, 2004). These are also critical for
DELTAS 239

FIG. 3.—Early example of a delta clinoform, showing topset, foreset, and bottomset strata (Scruton, 1960). A) Lithostratigraphic
representation shows facies boundaries as undulating but apparently sharp. Arrows indicate direction of progradation. Most
modern delta studies still show facies contacts in this manner. B) Correct representation of facies boundaries versus timelines.
Bed boundaries are more likely to follow the time lines (From Gani and Bhattacharya, 2005).

creation of traps in many shelf and offshore deltas such as the 1991; Jennette and Jones, 1995; Mellere and Steel, 1995, 1996;
Gulf of Mexico and Nigeria (e.g., Evamy et al., 1978; Berg and Dalrymple, 1999; Willis et al., 1999; Willis and Gabel, 2001;
Avery, 1995). Bhattacharya and Willis, 2001; Ta et al., 2002; Ta et al., 2005; Hori
A large body of research, begun with advent of seismic and et al., 2002; Davies et al., 2003; Allison et al., 2003; Dalrymple et al.,
sequence stratigraphy, emphasizes the evolution of modern and 2003; Lambiase et al., 2003; White et al., 2004; Willis, 2005). There
Quaternary deltas in the context of relative sea-level changes has also been an increasing focus on studies of muddy prodeltaic
(Dominguez et al., 1987; Boyd et al., 1989; Williams and Roberts, shelves linked to Modern highstand deltas, such as the Po,
1989; Carbonel and Moyes, 1987; Hart and Long, 1996; Hori et al., Ganges–Brahamaputra, Amazon, and Orinoco (Nittrouer et al.,
2002; Ta et al., 2002; Sydow and Roberts, 1994; various papers in 1986; Kuehl et al., 1997; Correggiari et al., 2001; Liu et al., 2002;
Sidi et al., 2003; various papers in Anderson and Fillon, 2004), and Warne et al., 2002; Cattaneo et al., 2003; Correggiari et al., 2005;
the application of these concepts to ancient deltas (Galloway, Neill and Allison, 2005).
1989a, 1989b; Bhattacharya and Walker, 1991; Martinsen, 1993;
Tesson et al., 1993; Bhattacharya, 1994; Gardner, 1995; Garrison Definitions
and van den Bergh, 1997; Plint, 2000; Bhattacharya and Willis,
2001; Garrison and van den Bergh, 2004). Deltas have been defined as “discrete shoreline protuber-
Several recent studies have documented examples of tide- ances formed where rivers enter oceans, semi-enclosed seas,
influenced deltas, which until recently have been the least well lakes or lagoons and supply sediments more rapidly than they
documented in the ancient record (e.g., Maguregui and Tyler, can be redistributed by basinal processes” (Elliott, 1986, p. 113).
240 JANOK P. BHATTACHARYA

FIG. 4.—Sandbody geometries of the six delta types of Coleman and Wright (1975) plotted on the river-, wave-, and tide-dominated
tripartite classification of Galloway (1975), from Bhattacharya and Walker (1992). Note that all sand bodies narrow and thicken
towards a point (fluvial) source. Also note similarity of tide-dominated isolith to river-dominated end member.

By this definition, all deltas are to some degree river-influenced. larger wave-influenced delta systems (Bhattacharya and Giosan,
Deltas are therefore fundamentally regressive in nature 2003). In particular, a river can act as a groyne, or hydraulic
(Dalrymple, 1999). barrier, that traps sediment carried in the longshore drift system
The term delta has also been applied to many ancient facies (Fig. 6; e.g., Dominguez, 1996; Rodriguez et al., 2000; Bhattacharya
successions or clastic wedges that show a marine to nonmarine and Giosan, 2003). Barrier islands can also form during the trans-
transition, or which contain a marine–fluvial or lacustrine–fluvial gression of a delta, such as the Chandeleur islands in the Gulf of
interface (Alexander, 1989), following the early definition of Mexico, which are the remnants of a now-abandoned delta lobe
Barrell (1912). Although a shoreline must be crossed in such a of the Mississippi delta (Fig. 7; Boyd and Penland, 1988). Where
transition or interface, the identification of the shoreline as spe- basinal processes redistribute sediment to the point that the
cifically deltaic usually requires good three-dimensional control fluvial source and delta morphology can no longer be recognized,
of facies patterns. This may consist of maps of lithofacies distribu- more general environmental terms such as paralic, strandplain,
tions showing a thickening and narrowing of the clastic succes- or coastal plain may be more preferable (Alexander, 1989).
sion toward the point of fluvial input, and the required seaward Deltas occur at a wide variety of scales ranging from conti-
protuberance of the shoreline (Fig. 4). nental-scale depositional systems, such as the modern Missis-
sippi delta (Fig. 8), with an area of about 28,500 km2, to compo-
Distinguishing Deltas from Other Depositional Systems nents of other depositional systems such as bayhead deltas
within estuarine or lagoonal systems. Many continental-scale
Much of the sediment in a delta is derived directly from the deltas, such as the Danube in Romania (Fig. 9) and the Missis-
river that feeds it, in contrast to estuaries, in which sediment is sippi, may contain smaller-scale crevasse deltas within larger-
derived both from the marine and the fluvial realm (Dalrymple, scale lobes, resulting in a complex and hierarchical facies archi-
1999). Estuaries have also been defined as fundamentally trans- tecture.
gressive depositional systems, in contrast to deltas, which are
regressive (Dalrymple et al., 1992). RIVER-MOUTH PROCESSES
In barrier-island systems, sediment is supplied alongshore
(Reinson, 1992). The terms ebb-tidal delta and flood-tidal delta A delta forms when a river of sediment-laden freshwater
have also been applied to sediment accumulations that form enters a standing body of water, loses its competence to carry
around tidal inlet channels in barrier–lagoon depositional sys- sediment, and deposits it. The theory of jets has been widely
tems (Reinson, 1992). Barrier islands may form components of applied to explain the dynamics of how river plumes interact
DELTAS 241

FIG. 5.—Tripartite classification of deltas, into river-, wave-, and tide- dominated end members (Galloway, 1975). Tide-dominated
end members are noted as being “estuarine”. This prompted Walker (1992) to abandon the concept of a tide-influenced delta. Also
note that the São Francisco and Brazos deltas are considered as type examples of wave-dominated end members.

with the body of water that they flow into (e.g., Bates, 1953; hyperpycnal and hypopycnal plumes at the same time (Nemec,
Wright, 1977; Orton and Reading, 1993; Nemec, 1995). The 1995; Kineke et al., 2000). Homopycnal conditions are the least
internal facies distribution and external morphology of a deltaic common, because only small density differences are required for
deposit depends upon (1) whether the river outflow is more a flow to become either hypopycnal or hyperpycnal.
dense (hyperpycnal), equally dense (homopycnal), or less dense Much of the active sand deposition occurs in a distributary-
(hypopycnal) than the standing body of water, (2) the interac- mouth bar (also referred to as a stream-mouth bar or a middle-
tion of the river plume with marine processes, which can in- ground bar). Mouth bars are a fundamental architectural element
clude waves, tides, storms, and ocean currents, and biogenic in modern deltas; they can coalesce to form complex bar assem-
reworking (Fig. 10), (3) the physical position of the delta in the blages, which in turn build regional-scale depositional lobes.
basin, such as the shelf edge, and (4) the degree to which river- Mouth bars scale broadly to the width of flow, although flow
derived sediments are reworked by marine processes. widths in distributary channels can vary both spatially and
Historically, most marine deltas have been assumed to be temporally. Individual bars can be on the order of several kilome-
hypopycnal, but many rivers experience dramatic changes in ters long in relatively large rivers like the modern Atchafalaya
discharge as a function of seasonal climate change or as a result delta (Fig. 11; Van Heerden and Roberts, 1988; Tye, 2003). The size
of major floods associated with storms. As a consequence, many and shape of a mouth bar also depends on the angle of dispersion
rivers can alternate from hypopycnal to hyperpycnal conditions, of a plume, the flow conditions (hyperpycnal, hypopycnal, or
even in fully marine settings (Nemec, 1995; Mulder and Syvitski, homopycnal), and the forces that act on the river plume (buoyant,
1995; Parsons et al., 2001). Many river plumes may show both inertial, and frictional forces and basinal processes).
242 JANOK P. BHATTACHARYA

Symmetric Asymmetric Deflected


A < 200 A > 200 A > 200

LEGEND
Net sediment drift at mouth River sediment discharge
Lagoonal Facies
Fluvial & Bayhead Delta Facies
Beach & Barrier Sand
Pre-delta

FIG. 6.—Morphology of wave-influenced deltas. Top row represents lower fluvial discharge compared to bottom row. River plume
acts as a groyne that traps sediment updrift (after Bhattacharya and Giosan, 2003). Asymmetry index represents the ratio of fluvial
sediment discharge to alongshore sediment transport rate.

FIG. 7.—Evolution of Mississippi delta lobes


from progradation to abandonment (from
Boyd et al., 1989). Delta goes through an
initial cycle of progradation, during which
it shows a river-dominated character. As
it is abandoned, it forms into a barrier-
lagoon system. The barrier is ultimately
drowned to form a relict shelf shoal.
DELTAS 243

tion described earlier, may initiate growth faults (Bhattacharya


and Davies, 2001, 2004) and may be important in causing avul-
sions as distributary channels become choked with sediment and
switch course.
Waves smooth out and elongate mouth bars in a shore-
parallel direction (Fig. 12; Wright, 1977; Fielding et al., 2005b).
The ability of waves to extend a bar downdrift depends on the
ratio of flood frequency to longshore-drift transport capacity. In
deltas with high-wave-energy regimes or very infrequent floods
(e.g., centennial floods), mouth bars may be extended for many
kilometers or more alongshore. Tides commonly dissect the bar,
or elongate it in a shore-normal direction. Tides may also cause
distributary channels, and in turn the associated bars, to be stable
for centuries, resulting in length-to-width ratios of up to 10
(Reynolds, 1999).

Hypopycnal (Buoyancy- and Friction-Dominated) Deltas

Where a river enters salt water, the density of the fresh river
water plus suspended sediment load may be less than that of the
sea water, causing hypopycnal flow (Fig. 10). Suspended muds
are carried out into the receiving basin as a buoyant plume,
resulting in lower depositional slopes.
Hypopycnal mud plumes may be deflected along the shelf by
waves, ocean gyres, or other oceanic circulation currents. In cases
where this mud is trapped within the littoral zone, it may form a
hyperconcentrated fluid mud that accretes to the shoreline. Win-
FIG. 8.—Infilling of interdistributary bays by historically dated nowing of this mud may cause shells or sands to form thin beach
crevasse “subdeltas” in the modern Mississippi birdfoot delta. deposits that armor the underlying mud and allow the downdrift
Note the large variation in scale of deltas and distributary muddy coastline to prograde, forming a chenier plain (Rine and
channels. At least three orders of branching can be discerned Ginsburg, 1985; Augustinius, 1989; Penland and Suter, 1989;
(from Bhattacharya and Walker, 1992; simplified from Coleman Draut et al., 2005). These are common on the downdrift margins
and Gagliano, 1964). of muddy delta systems, such as the chenier plain of the Louisiana
coast, which lies downdrift of the mighty Mississippi outflow
(Penland and Suter, 1989; Draut et al., 2005). The Camau penin-
Tye (2004) compiled data on the dimensions of modern mouth sula is a largely muddy accumulation that forms the downdrift
bars and other sandy elements in the Atchafalaya delta in the Gulf wing of the Mekong delta (Ta et al., 2002). Nearly 50% of mud
of Mexico and the Colville, Kuparuk, and Sagavinortik rivers in trapped in the modern Orinico delta in Venezuela is actually
Alaska. His data showed that bar widths ranged from 100 m to 3 derived from the Amazon (Rine and Ginsburg, 1985; Warne et al.,
km and bar lengths ranged from 140 m up to a maximum of nearly 2002). The Amazon muds are thus carried along the shelf for a
7 km. Modal mouth-bar widths are between 120 to 410 m, and distance of over 1000 km. Mud from the modern Po delta has also
modal lengths are between 250 m to 610 m. A compilation by been tracked for several hundred kilometers to the south along
Reynolds (1999) of ancient mouth-bar sand bodies shows consid- the Adriatic coast (Fig. 13; Cattaneo et al., 2003).
erably larger dimensions. His study showed that mouth-bar sand River bedload typically stops moving at the point of flow
bodies range from 1.1 km to 14 km wide with lengths of between expansion, forming the mouth bar, whereas suspended-load
2.6 km to 9.6 km. Average sand-body widths are about 3 km, and muds continue to be transported basinward. Hypopycnal deltas
average lengths are about 6 km. Reynolds suggests that mouth- are thus characterized by a distinct separation of the friction-
bar sandstones are typically twice as long as they are wide. inducing sandy bedload from the buoyant suspended muddy
Clearly, the average values of ancient sand bodies (Reynolds, load. Depending on plume stability, muddy plumes may pro-
1999) versus their modern geomorphic counterparts (Tye, 2003) duce subaqueous distributary channels with well developed
illustrate that ancient examples represent the migration and levees that may cause the mouth bar and channel to form elongate
growth of modern bars, and hence give larger dimensions. bar-finger sands (Fisk, 1961).
Because bars are bedload features, they induce an enormous At low stage, the more dense sea water can intrude many
amount of form friction, in excess of that associated with grain kilometers upstream into the river, forming a salt wedge, as is
and bedform roughness. This form friction significantly lowers seen in many modern deltas, such as the Po in Italy (Nelson, 1970).
bed shear stress and causes channel discharge to decrease as well Salt wedges can bring marine or brackish fauna into the distribu-
as causing a change in the direction of flow around the bar (i.e., tary channels, as observed in delta-plain distributary channels in
bifurcation of the channel). As bifurcation continues, the system the Cretaceous Ferron Sandstone in Utah, U.S.A. (e.g., Corbeanu
may become unstable, initiating an autocyclic upstream avulsion et al., 2004). During low discharge, muds in the overlying buoy-
of the feeding distributary. ant plume may flocculate and be deposited through suspension,
Mouth bars can accrete downstream, laterally, and upstream settling as extensive bar drapes in the river and mouth-bar areas.
(e.g., Van Heerden and Roberts, 1988; Corbeanu et al., 2004; These bar drapes can form fluid-mud layers that may be flushed
Olariu et al., 2005; Olariu and Bhattacharya, 2006). Downstream out onto the shelf during subsequent floods. Settling of sediment
accretion is an important process by which deltas grow and within a hypopycnal plume may result in unstable fingers of
prograde. Upstream accretion of sand, caused by the form fric- sediment collapsing through the water column and becoming
244 JANOK P. BHATTACHARYA

-
Chilia Arm

i
rien
4b

Jeb
4a 4c

Letea

TULCEA
Sulina Arm
Sf .
Gh
eor
ghe
Arm
2

Ca
ra
orm
an
1

ile
ur
t


Razelm
Lake 3a
3b
3
nd
la
Is
in
c al
Sa

SEA
CK
Sinoe
A
BL
Lake

Bedrock
Delta Plain
Abandoned Channels
Beach and Barrier Sands
0 3 6 9 12
km

FIG. 9.—History of the Danube delta plain. The highest-discharge, northern branch feeds a highly river-dominated delta lobe 4c,
comprising numerous bifurcating distributary channels with only minor wave reworking. The southernmost branch feeds the
distinctly asymmetric wave-influenced lobe 3. The updrift side of lobe 3 comprises amalgamated beach ridges of the Saraturile
Formation whereas the downdrift side comprises river-dominated bay-head deltas (3b) building behind a wave-formed barrier
island. The asymmetry is preserved in the older lobes 1 and 2. The central lobe 2 is largely inactive and is presently being destroyed.
Sands from lobe 2 are carried south by longshore drift to accumulate in the vicinity of lobe 3. Successive ages and outlines of lobes
are: 1, 9000–7300 yr BP; 2, 7300–2500 yr BP; 3 and 4, 2900 yr BP–present (based on radiocarbon dates of Panin et al., 1983). Figure
is based on map prepared by Gastescu (1992).
DELTAS 245

Hyperpycnal
Buoyancy-Dominated

Salt Wedge

Levee Bar Finger

Transverse Flow
Convergence

Low Tide
Friction Buoyancy

Friction Turbulent
Friction Waves Diffusion
Buoyancy
&

Bifurction Mixed-Influence Flows


around mouth bar High Tide
Inertia-Dominated
Homopycnal Flow

Gilbert Foresets

Hyperpycnal Flow

Loading/Invasion
Mass Flow Freezing
FIG. 10.—Examples of mouth-bar processes in river-dominated deltas (from Reading and Collinson, 1996, after Orton and Reading,
1993) incorporating ideas of Bates )1953), Wright (1977), and others. See text for discussion.
246 JANOK P. BHATTACHARYA

A B LOCATION
Ba
yo
u
Te
c he

EAST
PASS
Six Mile Lake

Morgan City

t
tl e

iv er
Ou
ISLE DERRIERE

ya R
ke
a
xL

fa la
Marsh Island

Wa
POULE

A tc ha
navigation A tc
D'EAUX ha
fa la
channel ISLANDS
ya
Ba
0 10 20 km y

NA LOUISIANA

1973 1975 TA Gu
L lf o
CH.
fM
e x ic o
Mis
sis
sip
pi
Riv
er

Atchafalaya Bay

CH. LEGEND
IA
TR
NU Exposed area
1973
1976
1km 1982
Mouth bar growth
1976 1979 Upstream
Downstream
Lateral
SOUTHE
AST PAS
S

0 500 1000 m
Scale

FIG. 11.—Development of a shallow-water delta in Atchafalaya Bay, Mississippi Delta, U.S.A. A) River-dominated lobe forms by
the coalescing of distributary-mouth bars (black), suggesting friction dominance. B) As the delta grows, the mouth bars accrete
upstream and downstream (compare 1976 and 1982 shorelines). Note that there are numerous orders of distributary channels,
culminating in small terminal distributary channels. Also note the scale of the mouth bars, which are on the order of several
hundred meters wide and one to several kilometers in length (from Olariu and Bhattacharya, 2006; after Van Heerden and
Roberts, 1988).

concentrated enough to produce a hyperpycnal flow (Nemec, been shown to occur in marine setting, at sediment concentra-
1995; Parsons et al., 2001). tions of 1–5 kg/m3 (Parsons et al., 2001). Such low-concentration
hyperpycnal flows may occur where marine water is colder
Hyperpycnal (Inertia-Dominated) Deltas than fresh river outflow, or where the shallow marine setting is
brackish, such as occurs at many delta fronts. Hyperpycnal
In freshwater lakes, sediment concentrations less than 1 kg/ flows may cause sediment to bypass the shoreline or mouth bar
m3 produce hyperpycnal conditions whereas sediment concen- and be deposited on the offshore shelf as density underflows.
trations greater than the density caused by dissolved salt in Because the momentum of the hyperpycnal flow exceeds the
seawater (about 35 to 45 kg/m3) may be required to generate ability of the standing body of water to stop the motion or lift the
hyperpycnal flows in marine settings (Mulder and Syvitski, plume by buoyant forces, hyperpycnal deltas have been re-
1995; Parsons et al., 2001). These flow conditions dominate ferred to as “inertia-dominated” (Bates, 1953). Hyperpycnal
where sediment-laden streams enter freshwater lakes, as occurs flows can be important in feeding deep-water systems, espe-
in many alpine or periglacial environments (Eyles and Eyles, cially during times of low sea level or in areas with narrow
1992). Many marine settings, however, are also hyperpycnal shelves where the river may be delivering sediment directly
(e.g., Wright et al., 1988; Mulder and Syvitski, 1995, Plink- into deep water. The resulting deposit of a hyperpycnal flow is
Björklund and Steel, 2004), and hyperpycnal conditions have either a fluid mud, or a silty or sandy graded bed (i.e., a
DELTAS 247

Symmetrical Deflected

FIG. 12.—A) Symmetrical mouth bars, versus B) deflected mouth bars. As a result of oblique wave approach (compare with
Figure 6). From Reading and Collinson (1996), after Wright (1977).

turbidite; Fig. 14A, B, C). Sands may occur as thinner, wedge- may also indicate more sustained flows (Plink-Björklund and
shaped sheets or fining-upward shallow undulating channel Steel, 2004).
deposits (Olariu et al., 2005; Gani and Bhattacharya, 2005; Plink- For sandy systems there has been significant debate about
Björklund and Steel, 2004, 2005). Hyperpycnal turbidites typi- how sandy delta front turbidites form. One hypothesis is that
cally show more complex internal geometry than surge-type delta-front-turbidites are fed by true hyperpycnal flows directly
turbidites (Mulder and Alexander, 2001; Plink-Björklund and from the proximal delta front caused by rapid sedimentation
Steel, 2004). The sustained flows associated with hyperpycnal (e.g., Mulder et al., 1996). The alternate hypothesis is that sandy
flows may result in thick, massive beds that typically show sediment is first “stored” in a proximal mouth bar, which builds
inverse grading at the base (associated with increasing flood up to a threshold slope and then becomes unstable. Floods,
discharge) followed by normal grading as the flood wanes. storms, or earthquakes may trigger a delta-front sediment grav-
Alternation of structureless to parallel-laminated sandstones ity flow. Both processes have been documented for the Modern

FIG. 13.—Mud from the Po delta , deposited during the Holocene highstand (HST) is carried
several hundred kilometers along the Adriatic coast by geostrophic currents (from
Correggiari et al., 2001).
248 JANOK P. BHATTACHARYA

FIG. 14.—A) Aggrading wave-rippled sandstones interbedded with normally graded siltstones and claystones, Cretaceous Dunvegan
Formation, Alberta, Canada. B) Interbedded normally graded very fine-grained sandstones and siltstones with lightly burrowed
mudstones. Prodelta mudstones of the Cretaceous Dunvegan Formation, Alberta, Canada. C) Normally graded to flat-stratified
sandstones of the Cretaceous Panther Tongue sandstone, Utah, U.S.A. A–C are interpreted as delta-front sediment-gravity-flow
deposits. D) Pervasively bioturbated, non-deltaic sandy mudstone, Cretaceous Dunvegan Formation, Alberta. Except for
hammer in Part C, scale bar is 3 cm.
DELTAS 249

Sepik river mouth in Papua New Guinea (Kineke et al., 2000). In examples of marine steep-fronted Gilbert type deltas, such as the
the second scenario, all that is required is rapid sedimentation of Modern Alta delta in Norway (Corner et al., 1990). Numerous
the sandy load of the river, which could occur in either hypopycnal ancient examples are given in Colella and Prior (1990), and more
or homopycnal settings. In the second scenario, mud may never recently published examples from Europe include Burns et al.
be deposited from a hyperpycnal flow, and the deposits thus (1997), Ulicny (2001), and Soria et al. (2003).
comprise mud deposited from hypopycnal flows that shows less
“fluvial” influence (e.g., more normal marine biota), interbedded DELTA ENVIRONMENTS
with rapidly deposited sands with burrowed tops (MacEachern
et al., 2005). Muds deposited from suspension accumulate at rates Deltas comprise three main geomorphic environments of
an order of magnitude slower than hyperpycnal muds and con- deposition (Fig. 15): the subaerial delta plain (where river pro-
sequently show much higher degrees of bioturbation (Fig. 14D; cesses dominate), the delta front (the coarser-grained area where
Allison et al., 2000). river and basinal processes interact), and the prodelta (primarily
Rivers that frequently experience hyperpycnal conditions are muddy). These three environments roughly coincide with the
typically small “dirty” systems that drain high-relief, tectonically topset, foreset, and bottomset strata of early workers, although
active terrains, such as the Eel River, which feeds the Northern the boundaries overlap and specific definitions of the delta front
California coast (Mulder and Syvitski, 1995; Syvitski and are not widely agreed on.
Morehead, 1999). However, these systems are usually not
hyperpycnal throughout the year. Most sediment discharge oc- Delta Plain
curs during rare, large-magnitude floods. Ninety percent of the
yearly Eel River discharge, for example, occurred in just a few The delta plain is defined by the presence of distributary
days of flooding. The rest of the year, the Eel is hypopycnal and channels. It includes a wide variety of nonmarine to brackish,
carries very little sediment (Syvitski and Morehead, 1999). Dur- paralic to wetland sub-environments including swamps, marshes,
ing low-discharge periods, sediments deposited during major tidal flats, lagoons, and interdistributary bays. Although readily
floods may be significantly reworked by waves and tides. distinguished in most modern delta environments (e.g., Ta et al.,
A “pulsed” depositional history characterizes many deltas, 2005; Fielding et al., 2005a), the distinction of these various
such as the Brazos, in the Texas Gulf Coast, the Danube, in the shallow-water, brackish wetland environments is not routinely
Black Sea, the Senegal in Africa, and the Burdekin delta, in attempted in ancient settings (but see McCabe, this volume).
Australia (Bhattacharya and Giosan, 2003; Fielding et al., 2005a). The landward limit of modern delta plains is typically taken
Growth of these deltas is confined to very short periods of major at the point in the alluvial realm where trunk streams become
flood activity associated with storms. Depending on flood fre- unconfined and distributive (typically immediately downstream
quency, which can be seasonal or centennial, flood-borne sedi- of the alluvial valley). In many cases, this is the nodal avulsion
ment can be completely reworked over time. The Brazos and point on an alluvial plain. In modern settings, the delta plain can
Burdekin deltas have recently been redefined as “flood-domi- be subdivided into a lower delta plain, marked by tidal incur-
nated” rather than wave-dominated systems (Rodriguez et al., sion of sea water, and a more landward upper delta plain, in
2000; Fielding et al., 2005a, 2005b). Storms at the downstream end which major distributary channels still occur but in which there
of a delta system may do little to increase sediment discharge, is no incursion of marine water (Fig. 15; Coleman and Prior,
even though they may cause flooding, whereas storms in the 1982).
hinterland may be far more important in terms of increasing The demarcation between these areas is referred to as the bay
sediment discharge. In continental-scale deltas, such as the Nile, line (Posamentier et al., 1988). Rivers may experience tidal modu-
the Amazon, or the Mississippi, there may be no obvious link lation of flow far upstream of the actual marine incursion, de-
between coastal storms and hinterland storms, which may be pending on the ratio of tidal forces to river discharge. The land-
completely out of phase. Rivers associated with continental drain- ward limit of incursion depends on slope and discharge. In
ages, in excess of 106 km2, have been suggested to rarely, if ever, ancient settings, the bay line may be indicated by the landward
go hyperpycnal (Mulder and Syvitski, 1995), although if marine limit of marine or brackish-tolerant fossils or trace fossils, al-
waters are already brackish, or where marine water is cold, even though tidally influenced cross stratification may occur farther
large rivers may go hyperpycnal frequently (Parsons et al., 2001; upstream of any measurable brackish influence. The seaward
Plink-Björklund and Steel, 2004). limit of the lower subaerial delta plain is defined at either the
high-tide shoreline (Elliott, 1986) or the low-tide shoreline, which
Homopycnal (Friction-Dominated) Deltas includes the foreshore (e.g., Coleman and Prior, 1982).
The upper delta plain is a fluvial environment, although in
In homopycnal settings there may be a greater degree of rare cases it may be indirectly tide influenced. Lakes lack tides,
mixing between the river and standing body of water. These and consequently the distinction between the upper and lower
situations are common in fresh-water deltas and can also occur in delta plain cannot be made in lacustrine deltas. Steeply sloping
marine settings where the amount of bed load is high. In shallow fan deltas, adjacent to scarps, have very limited delta plains.
water, friction at the bed causes rapid deceleration and develop-
ment of a mouth bar that causes the associated distributary Delta Front
channel to bifurcate, and settings of this kind have been referred
to as “friction-dominated” (Wright, 1977). However, friction is The delta front is defined as the shoreline and adjacent
important in both hypopycnal and hyperpycnal deltas. “Friction- dipping sea bed (Elliott, 1986). It is defined as the area domi-
dominated” mouth bars are more fan shaped than buoyancy- nated by coarser sediment (sand or gravel) that includes sub-
dominated mouth bars, and may be dominated by traction- aqueous topset and foreset beds. However, many studies of
current features such as climbing ripples and cross bedding (Fig. modern delta systems do not include the foreshore and shoreface
10). Deltas of this kind are characterized by close-to-angle-of- environments within the delta front but rather treat this as a
repose foreset beds, such as seen in the gravelly freshwater deltas separate intertidal to subtidal “delta platform” environment
originally described by Gilbert (1885). There are several good (Fig. 16; e.g., Coleman and Prior, 1982; Ta et al., 2002; Roberts
250 JANOK P. BHATTACHARYA

tions of distributary channels may be fixed for long periods,


forming elongate bar fingers, as in the deeper-water mud-domi-
nated Mississippi “birdfoot” delta (Figs. 8, 17; Fisk, 1961). How-
ever, the elongation of the modern Mississippi “birdfoot” delta is
somewhat artificial, because it has been maintained for many
decades by the U.S. Army Corps of Engineers. By contrast, in
siltier or sandier systems deposited in shallower water, or not
stabilized by human interference, distributaries switch more
rapidly and coalesce to form more lobate deltas, as in the Lafourche
(Fig. 17) and Atchafalaya (Fig. 11) deltas (Olariu and Bhattacharya,
2006).
The seaward-dipping slope associated with the distal mar-
gin of a distributary-mouth bar is also sometimes referred to as
the distal delta front and can form a relatively continuous
sandy fringe in front of the active zone of mouth bars. Internally,
the distal bar is physically built by rapidly decelerating “frontal
splays”. In high-slope delta fronts, these can be expressed as
normally or inversely graded beds deposited from waning
turbidites or grain flows (Por™bski and Steel, 2003; Plink-
Björklund and Steel, 2005; Olariu et al., 2005) In coarser deltas,
these deposits produce the classical foreset geometries that
define Gilbert deltas (e.g., Soria et al., 2003). Several researchers
(e.g., Coleman and Prior, 1982) reserve the term delta front to
refer only to this distal bar environment. The term delta front has
also been applied to refer to mid-shelf muddy clinoform strata,
seaward of any significant sand deposits (Fig. 16; Roberts and
Sydow, 2003).
FIG. 15.—Major areal subdivisions of a delta. The upper delta
plain is essentially nonmarine and characterized by distribu- Delta Front Versus Shoreface
tive river systems.
In wave-influenced deltas, mouth bars may be reworked into
a distinct shoreface, which can be considered as part the delta
and Sydow, 2003). The width of this subtidal platform can be up front (Barrell, 1912). The shoreface is the seaward-dipping equi-
to several kilometers where tidal range is high (e.g., Corner et librium surface that forms in response to the asymmetry of
al., 1990; Hori et al., 2002; Ta et al., 2002; Allison et al., 2003; shoaling fair-weather waves (Barrell, 1912; Bruun, 1962; Swift,
Roberts and Sydow, 2003). 1968). The shape and extent of the shoreface depends on the
River-dominated delta fronts typically consist of a complex interaction of sediment supply and wave energy expended at the
association of terminal distributary channels and mouth bars that coast (Walker and Plint, 1992). Non-oceanic settings, or coasts
coalesce to form bar assemblages and depositional lobes (e.g., with wide shelves, typically lack swell waves, for example, and
Van Heerden and Roberts, 1988). In hypopycnal river-dominated have a correspondingly diminished shoreface in which fair-
settings, especially those with minimal tides or waves, the posi- weather waves affect only sediments deposited in a few meters or

FIG. 16.—Morphometric subdivisions of the Mahakam delta, Kalimantan, Indonesia. Note that muddy subaqueous foreset is referred
to as the “delta front” (modified after Roberts and Sydow, 2003).
DELTAS 251

FIG. 17.—Representative modern examples of river-dominated, wave-dominated, and tide-influenced deltas. Modified from Fisher
et al. (1969). River-dominated deltas are classified into lobate (shoal-water) and elongate (deep-water or birdfoot) deltas. In the
Mahakham example (after Allen et al., 1979), delta-front deposits comprise sandy siltstones and mudstones. Figure from
Bhattacharya and Walker (1992).

less. This may nevertheless impart a smooth-fronted appearance Prodelta


to the delta, but sediments deposited below the effects of fair-
weather waves will record the original depositional processes. In The prodelta has historically been interpreted as the area
mesotidal or macrotidal settings, tidal process may mask the where fine mud and silt settle slowly out of suspension. Prodelta
effects of waves. Many smooth-fronted modern deltas, such as deposits may be more or less burrowed, depending on sedimen-
the Brazos, Burdekin, Baram/Trusan, and Mekong, while show- tation rates. Prodelta muds may merge seaward with fine-grained
ing the effects of shallow-wave reworking, show a dominance of hemipelagic and commonly calcareous sediment of the basin
river-flood or tidal facies in the underlying sediments (Rodriguez floor. The preservation of silty or sandy lamination is commonly
et al., 2000; Lambiase et al., 2003; Ta et al., 2002, 2005; Fielding et taken to mark the influence of the river, as opposed to total
al., 2005a, 2005b). bioturbation of the basin-floor sediments in areas away from the
In wave-modified deltas, the shoreface may form part of the active river (Fig. 14D; Allison et al., 2000; Neill and Allison, 2005;
delta front. Shorefaces can also form in the absence of a river. The MacEachern et al., 2005). Where the sediments are rhythmically
shoreface can also be entirely erosional, especially during trans- laminated, a tidal influence may be inferred (Smith et al., 1990).
gression, where sediment supply may be minimal (e.g., Bruun, Because of the abundant suspended sediment, certain types of
1962; Swift, 1968; Nummedal and Swift, 1987; Kraft et al., 1987). vertical filter feeders and other organisms that produce open
This erosion forms a ravinement surface that commonly removes vertical burrows of the Skolithos ichnofacies tend to be suppressed
5–10 m of the topset portions of a delta. These ravinement (e.g., Moslow and Pemberton, 1988; Gingras et al., 1998;
surfaces form profoundly significant bounding discontinuities MacEachern et al., 2005).
that are the key to identifying and mapping ancient top-truncated The term prodelta and shelf have been presented historically
delta deposits (e.g., Weise, 1980; Walker and Plint, 1992; Hart and as mutually exclusive environments (e.g., Walker, 1984), which
Long, 1996; Posamentier and Allen, 1999; Bhattacharya and Willis, may be a serious error. Many of the world’s muddy shelves, such
2001; Martinsen, 2003). as the Adriatic (Fig. 13), Black Sea, Amazon, Bay of Bengal, Papua
252 JANOK P. BHATTACHARYA

A B

50 km 50 km

C D

10 km

E
10 km

20 km

FIG. 18.—Comparison of distributary-channel branching patterns in a river-dominated versus wave-dominated deltaic coastline. A)
River-dominated Lena River delta (Russian Arctic) shows numerous orders of branching with many tens of terminal distributary
channels. B) Wave-dominated coastline associated with the Paraíba do Sul, Brazilian coast. C) Po delta, Italy. D) Ebro delta, Spain.
Bifurcation is inhibited in wave-dominated deltas because the river is unable to prograde into the basin as rapidly. This effectively
allows the river to maintain its grade, which in turn inhibits avulsion. E) Tide-dominated Ganges–Brahmaputra delta shows
highly elongate channels. Photos courtesy of NASA.
DELTAS 253

New Guinea, Gulf of Mexico, and others, are now being inter- higher slope, which inhibits avulsion. As a consequence, wave-
preted as the subaqueous extension of deltas (e.g., Nittrouer et al., influenced deltas typically have only a few active distributary
1986; Kuehl et al., 1997; Michels et al., 1998; Liu et al., 2002; channels (Figs. 18B, C, D) whereas river-dominated deltas can
Cattaneo et al., 2003; Roberts and Sydow, 2003; Kuehl et al., 2005; have tens to hundreds of active terminal distributary channels
Neill and Allison, 2005). Studies of modern muddy shelves show (Fig. 18A).
that much of the muddy sediment deposited by suspension out Many tidally influenced deltas show distributary channels
of buoyant river plumes ultimately concentrates at the seabed, that are stable for hundreds to thousands of years, as in the
forming a fluid-mud layer that may be kept in suspension by Mekong delta (Ta et al., 2002; Ta et al., 2005) and Ganges–
waves (e.g., Kineke et al., 1996) or moved by storms (Allison et al., Brahmaputra (Kuehl et al., 2005). This results in the develop-
2000; Draut et al., 2005). Mud may also be introduced directly ment of elongate bars and islands that can be tens of kilometers
onto the seafloor by hyperpycnal flows (Mulder and Syvitski, in length and a few kilometers wide (Fig. 18E). Increased chan-
1995). Because sedimentation rates of fluid muds are so much nel stability results in far more elongate sand bodies, with
higher than by suspension settling, they are probably far more higher length-to-width ratios than are typically found in river-
important in the construction of the shelf than has historically dominated delta fronts (ratio of 10 versus 2, respectively;
been realized (Neill and Allison, 2005). Fluid muds may be Reynolds, 1999).
characterized by centimeter- to decimeter-thick beds which show In systems with many orders of branching, younger and
a lack of lamination and bioturbation and which are interbedded shorter-lived distributary channels, lower on the delta plain, tend
with graded siltstone or sandstone beds (Kuehl et al., 1986; to be straighter because of lower slopes and lower discharge,
Allison et al., 2000; MacEachern et al., 2005). whereas longer-lived channels on the upper delta plain can show
complex and highly sinuous or braided channel patterns because
Distributary Channels of higher slopes. In wave- or tide-dominated deltas, rivers may
retain their trunk character all the way to the shoreline. This is
Distributary channels may show a wide range of sizes and why the Ganges–Brahmaputra and São Francisco rivers are
shapes in different positions on the delta (Fig. 18; Olariu and braided at the river mouth.
Bhattacharya, 2006). There is therefore no such thing as “a Terminal distributary channels are difficult to recognize in
distributary channel” in many deltas. Typically, a trunk fluvial subsurface because they tend to be shallow. Successive higher-
system first avulses at the point where the river becomes uncon- order distributary channels should become thinner and narrower
fined, forming a nodal avulsion point (e.g., Nelson, 1970; Mackey downstream, as has been documented in ancient channels within
and Bridge, 1995). Delta-plain channels tend to be few in num- the Devonian “Catskill” deltaic wedge of the Appalachian basin,
ber and are separated by wide areas of interdistributary bays, U.S.A. (Bridge, 2000). Thus, one way to determine if an ancient
swamps, marshes, or lakes on the delta plain, although these fluvial system is distributive is to see if the widths and depths of
interdistributary areas can be replaced by channel deposits, channels become smaller in more distal reaches of a clastic wedge.
depending on the avulsion frequency and rate of channel migra-
tion (Bristow and Best, 1993; Mackey and Bridge, 1995; Holbrook, Distributary Channels Versus Incised Valleys, and
1996). In an ancient setting, upper-delta-plain channels may be Overapplication of the Mississippi Analogue.—
very difficult to distinguish from fluvial channels, especially if
there is no tidal influence. Many ancient examples of so-called river-dominated deltas
Distributary channels can show several orders of branching, exhibit thick channelized deposits overlying marine prodelta
readily measured in modern systems but difficult to determine in shales and have been interpreted as distributary channels cutting
ancient examples. The smallest-scale channels are referred to as into their associated delta fronts (Fig. 19; e.g., Busch, 1971, 1974;
“terminal distributary channels” and are intimately associated Cleaves and Broussard, 1980; Rasmussen et al., 1985) despite the
with mouth bars that form at the distal delta plain and proximal fact that this is rare in many modern deltas. Some of these sand-
delta front (Olariu and Bhattacharya, 2006). Terminal distribu- stones are over 30 m thick and cut out delta-front sandstones that
tary channels can extend several kilometers offshore, forming are only 10 m thick. Many of these deeply incised channels are now
channelized to scoured facies within the delta front (Olariu and recognized as valley fills rather than distributary channels (Willis,
Bhattacharya, 2005). 1997; Bowen and Weimer, 2003; Bhattacharya and Tye, 2004).
Distributary-channel bifurcation occurs at a point where the The earlier interpretations of these features as distributary
channel can no longer cut directly through the distributary- channels stemmed largely from comparison with the deep and
mouth bar, forcing it to split into two smaller channels flanking stable distributary channels of the modern Mississippi birdfoot
the bar crest. Channel-bifurcation frequency and branching pat- delta. The fact that deep Mississippi distributary channels erode
terns are strongly dependent on slope, river discharge, water into underlying prodelta muds has been cited as the main reason
depth, and the interaction of the river plume with marine pro- why distributary channels, in general, do not migrate laterally
cesses (Fig. 18). Multiple bifurcations are favored in low-gradi- (Coleman and Prior, 1982). However, the Mississippi is a conti-
ent, high-discharge, river-dominated deltas, where friction is the nental-scale system, and the modern channel has been kept in
dominant process controlling sediment dispersal and deposition place through the dredging efforts of the U.S. Army Corps of
(Welder, 1959; Wright, 1977). Nodal avulsion of trunk streams Engineers. It is an inappropriate analogy with which to interpret
and distributary crevassing are common processes in river-domi- many shallower-water, mid-continent delta systems, such as
nated deltas because hydraulic gradients decrease as rivers and developed in the Pennsylvanian and Cretaceous systems of North
distributaries extend their courses. Friction, caused by mouth America, which drained considerably smaller areas (Bhattacharya
bars and plume dispersion, also reduces the discharge. and Tye, 2004).
In wave-modified deltas, much of the sediment delivered to
the shoreline is carried away from the river mouth by longshore Regional Controls on Delta Morphology
transport (Figs. 6, 18). Thus, relative to river-dominated deltas,
the progradation rate of wave-influenced deltas is slowed. This Many other factors may also influence the delta form, apart
allows rivers feeding wave-influenced coasts to maintain a from the nature of the fluvial input and the reworking of the
254 JANOK P. BHATTACHARYA

BOOCH
DELTA

McAlester Fm.

FIG. 19.—Ancient examples of dendritic “shoestring” sandstones of the Pennsylvanian Booch sandstone (Oklahoma). A) Map view
suggests river-dominated, elongate deltas. This interpretation was heavily biased by modern Mississippi birdfoot delta (after
Busch, 1971). B) Well-log cross section suggests over-thick valleys. Many of these systems should probably be reinterpreted as
incised valleys.

deposits by waves and tides. Coleman and Wright (1975) empha- nated, because they lie within major straits or continental reen-
size the nature of the receiving basin, the nature of the drainage trants, such as the tide-influenced Ganges–Brahmaputra delta
basin, the tectonic setting, and climate. In addition, relative sea- (Fig. 18E), which lies at the head of the Bay of Bengal. The tropical
level changes influence the extent of delta growth and destruc- deltas of southeast Asia lie within a part of the world that has very
tion. These factors are not all independent. Sediment type and low wave heights but high tides, which also contributes to the
rate of supply, for example, are a function of the size, relief, generally tide-dominated nature of deltas there (Nummedal et al.,
climate, and underlying geology in the drainage basin. Relief 2003). Sediment type and rate of supply were also influenced by the
may be dependent on the tectonics of the drainage basin. Wave or absence of land plants in pre-Devonian rocks, resulting in higher
tide energy may be a function of eustasy, shelf slope, and size and sedimentation rates and a greater proportion of fan deltas (Stow,
shape of the receiving basin, and wave energy is also related to 1986; MacNaughton et al., 1997). Deltas formed against scarps or
climate (e.g., wind direction and strength). faults typically form as fan deltas with little to no delta plain.
For example, tidal range is typically enhanced within coastal The focus of geologists on studying sandstone has resulted in
embayments. Many of the world’s largest deltas are tide-domi- the erroneous notion that many rivers carry primarily sand. Most
DELTAS 255

rivers carry between 85% to 95% mud (Schumm, 1972), chiefly in within discrete lobes (Fig. 9). This can create problems for
suspension. Mud-free rivers are rare in nature, and most modern interpreters, especially in subsurface studies, where the nature
mud-free rivers owe their lack of suspended material to the fact of a depositional system is typically determined on the basis of
that the suspended load is deposited in dams, far upstream of the sparse core data, which may not represent the whole system.
river mouth. There are very few studies of Modern systems that Furthermore, the terms “dominated” versus “influenced” have
document how mud is partitioned between the delta plain (i.e., never been adequately distinguished. These could be quantified
floodplain) versus the prodelta shelf environment and what are in terms of wave or tide energy expended at the coast versus
the key controlling factors on sediment partitioning (e.g., sediment discharge or sedimentation rate at the river mouth or
Goodbred and Kuehl, 1999). In contrast, sequence stratigraphic other measurable parameters (e.g., as attempted by Coleman
studies of ancient systems have demonstrated that sediment is and Wright, 1975, Hayes, 1979, and Dalrymple, 1992), but these
partitioned over geological time as a consequence of changes in are virtually impossible to measure or determine in an ancient
accommodation, sediment supply, and sea-level change (e.g., system (see, however, Bhattacharya and Tye, 2004). Another
Jervey, 1988; Posamentier et al., 1988; Helland-Hansen and approach would be to measure the physical proportion of facies
Gjelberg, 1994; Posamentier and Allen, 1999). that were formed by wave, tide, or fluvial processes. This
approach is more applicable to ancient systems, and it may be
CLASSIFICATION OF DELTAS especially applicable to reservoir or aquifer modeling, where
facies architecture may have a first-order control on flow behav-
The commonly used tripartite classification of deltas (Fig. 5; ior. In a study of the Baram and Trusan deltas of Borneo,
Galloway, 1975) is based on the idea that the ratio of fluvial, Lambiase et al. (2003) showed that despite a smooth-fronted
wave, and tidal processes results in different and identifiable external geometry, suggestive of wave domination, the internal
plan-view morphology of resulting deposits as well as charac- facies show a strong tidal signature. The plan-view shape of the
teristic internal facies successions. Unfortunately, there has modern Brazos delta has long been cited as a classic example of
been a natural tendency for workers to force-fit their particular a wave-dominated delta, but recent coring studies show a
example into one of the end-member categories (e.g., dominance of river-flood deposits and have led Rodriguez et al.
Bhattacharya and Walker, 1992; Dominguez, 1996), despite the (2000) to reclassify the Brazos as a river-flood-dominated, wave-
fact that most deltas are likely to be mixed-influence and plot influenced delta. Similar work on the Burdekin delta in Austra-
somewhere within the triangle. Many modern deltas, such as lia shows that despite its smooth-fronted external appearance,
the Danube, show a mixture of delta types both between and which historically led it to be classified as wave-dominated,

FIG. 20.—Tide dominated deltas. A) Tidally reworked mouth bars are highly elongated. Central part of delta is predominantly sandy
bars whereas mud is partitioned along sides of system (from Dalrymple, 1992). B) During transgression of the East China Sea,
mouth bars are reworked into elongate, shore-normal shelf ridges (after Yang and Sun, 1988).
256 JANOK P. BHATTACHARYA

FIG. 21.—Delta triangle of Galloway (1975) is extended to include sediment caliber as a fundamental control (from Reading and
Collinson, 1996, after Orton and Reading, 1993).

internal facies from cores and outcrops show a predominance of The term “braid delta” or “braidplain delta” has been used to
river-flood deposits (Fielding et al., 2005a, 2005b) refer to a sandy or gravelly delta front fed by a braided river and
River-dominated deltas display an overall digitate or lobate characterized by a fringe of active mouth bars (e.g., McPherson et
morphology (Figs. 4, 5, 8, 11, 17, 18). In contrast, wave-influ- al., 1987). This term must be used with caution. There are many
enced deltas show smooth-fronted lobes with arcuate to cuspate examples of braided rivers that feed highly wave-influenced
margins (Figs. 1, 4, 5, 6, 9, 17, 18). Tidal processes form sand deltas (e.g., the São Francisco and Paraiba do Sul in Brazil) or tide-
bodies oriented parallel to the directions of the tidal currents influenced deltas (e.g., the Ganges–Brahmaputra) that bear little
(Figs. 4, 5, 17, 18E, 20), typically perpendicular to regional resemblance to the so-called “braid deltas” of McPherson et al.
shorelines (Dalrymple, 1992; Maguregui and Tyler, 1991; Willis (1987).
and Gabel, 2001). The area of the marine-influenced lower delta Dalrymple et al. (1992) extended the delta triangle of Gallo-
plain and delta front may be quite extensive where tidal range way (1975) into three dimensions to include the sequence strati-
is large. In tide-dominated settings, mud is partitioned along graphic concept that the abundance of different depositional
the margins of the system (Fig. 20), in contrast to wave-domi- systems is a function of relative sea-level change (Fig. 23). They
nated estuaries, where a wave-formed barrier–beach complex emphasize the relationship between “regressive” delta-type sys-
protects the estuary mouth and mud accumulates in the lagoon tems and “transgressive” depositional systems, such as estuaries
behind the barrier. and barrier-lagoons. Although this is a valuable extension of
Coleman and Wright (1975) recognized that the geometry of Galloway’s work, missing from this diagram is the fact that many
a delta sand body should reflect the relative importance of fluvial deltas contain barrier-island–lagoon systems, tidal flats, and
and marine processes. All of their delta geometries (Fig. 4) em- even drowned abandoned distributaries, which may exhibit a
phasize narrowing and thickening of sands towards a point (i.e., strongly estuarine-type fill (Fig. 9).
fluvial) source, but the seaward margins differ, as explained Seismic-stratigraphic studies and sequence-stratigraphic stud-
above. Orton and Reading (1993) extended the Galloway classi- ies of deltas led to a recognition that depositional systems change
fication to include sediment type (Fig. 21). Postma (1990) pre- their character as a function of their physical and temporal
sented an independent classification scheme based on the type of position. For example, shelf-edge deltas tend to form at sea level
feeder system, water depth, and mouth-bar process (Fig. 22). This lowstands (Fig. 24; e.g., Edwards, 1981; Posamentier et al., 1992;
classification scheme does not, however, include waves or tides Tesson et al., 1993). Deltas deposited at the shelf edge are com-
as key parameters. monly unstable and develop impressive growth faults. Sand
DELTAS 257

FIG. 22.—Classification of coarse-grained delta types incorporating type of feeder system, water depth, and type of mouth-bar process
(from Reading and Collinson, 1996; after Postma, 1990).

bodies in these settings are often aligned along strike (Fig. 25), but Wright, 1975). Thick mudstone deposits that do not display a
this elongation is controlled by subsidence along the growth distinctive upward-coarsening facies succession may occur in
faults rather than by wave processes. areas away from the river mouth, such as where there is signifi-
cant alongshore diversion of a muddy river plume.
VERTICAL FACIES SUCCESSIONS Depending on the shoreline trajectory during progradation,
significant delta-plain facies can accumulate above and behind
Coleman and Wright (1975) presented, in addition to sand- the migrating and subsiding delta front (Fig. 27). Thicknesses of
body shapes, a series of composite vertical facies successions upward-coarsening facies successions may range from a few
through the prodelta, delta-front, and delta-plain environments meters to a hundred meters, depending on the scale of the delta,
of each of their delta types. These idealized facies successions the water depth, the shoreline trajectory, and the subsidence rate.
represent “norms” and may be very useful points of reference Delta-front sands may subsequently be partially eroded as the
for outcrop studies where three-dimensional control may be distributary channel migrates over its own mouth bar (Fig. 26).
limited. Although a single vertical profile may not be represen- Recent studies of deltas that build into very shallow water (e.g.,
tative, especially in deltas that show extreme lateral facies Kroonenberg et al., 1997; Overeem et al., 2003; Fielding et al.,
variability, 1D vertical profiles remain the most common data 2005a, 2005b) show that upward-coarsening successions are more
type on which any study of depositional systems, be it core, well difficult to produce, because there is little space to accumulate a
log, or outcrop data, is based. Lateral facies relationships, in thick underlying prodelta platform over which the delta can
contrast, can be interpreted directly from seismic data (e.g., build. Where channel flow depths are on the same scale as water
papers in Anderson and Fillon, 2004) or in continuous cliff depth, the channel flows more easily cannibalize underlying
panoramas (e.g., Willis et al., 1999; Willis and Gabel, 2001; Soria muddy facies and the facies succession is dominated by sharp-
et al., 2003; Barton, 2004; Garrison and van den Bergh, 2004). based coarsening-upward mouth bars or fining-upward dis-
Typical facies successions through the dominantly marine (pro- tributary-channel fills (e.g., Holbrook, 1996). These sharp-based
delta and delta front) and dominantly nonmarine (delta plain) mouth bars may produce facies succession that are very similar to
parts of deltas, mostly in river- and wave-dominated settings, those produced in other low-accommodation settings, such as
are outlined below. during forced regressions (Plint, 1988; Posamentier et al., 1992;
Fielding et al., 2005a, 2005b).
Prodelta and Delta-Front Successions In shelf-edge deltas (Fig. 24), thick upward-coarsening delta-
front successions can be preserved within the hanging wall of
Progradation of a delta commonly produces a coarsening- growth faults, although they show increasing dip with depth. In
upward facies succession (Fig. 26) showing a transition from the shallower landward portions, greater reworking by shallow-
muddier facies of the prodelta into the sandier facies of the delta- marine processes can result in more complex facies successions
front and mouth-bar environments (Elliott, 1986; Coleman and (Winker and Edwards, 1983).
258 JANOK P. BHATTACHARYA

FIG. 24.—Block diagram contrasting lobate shoal-water (or shelf-


phase) deltas and shelf-edge deltas. Note thickening of facies
across growth faults in the shelf-edge delta (From Bhattacharya
and Walker, 1992; after Edwards, 1981).

Neill and Allison, 2005). Wave-formed structures may occur at


the tops of graded sandstone beds, but they are less abundant
than in a more wave-influenced setting. If floods occur during
major coastal storms, sets of highly aggrading wave-rippled
sandstone beds may occur (Fig. 14A) and hummocky cross
stratification may be abundant. Soft-sediment deformation fea-
tures result from high sedimentation rates and are common in
river-dominated deltas (Fig. 29). Deposition of overpressured
prodelta muds may cause remobilization of the overlying delta-
front sand (Figs. 29, 30; Coleman et al., 1983; Bhattacharya and
Davies, 2001; Wignall and Best, 2004). Cores from prodelta and
delta-front deposits of the Mississippi show a complex facies
architecture dominated by upward-coarsening facies succes-
FIG. 23.—Delta triangle of Galloway (1975) as extended by
sions (Fig. 31).
Dalrymple et al. (1992) to reflect changes in sediment supply
Sandy delta-front facies predominantly reflect deposition
(from Reading and Collinson, 1996).
from rapidly decelerating unidirectional flows in distributary-
mouth-bar environments. Structures may include unidirectional
The specific nature of the facies and beds in prograding current ripples and cross bedding, flat-stratified sandstones, or
prodelta and delta-front successions depend on the processes massive graded beds (Fig. 14B, C), depending on the impor-
influencing sediment transport, deposition, and reworking. tance of frictional versus inertial processes (Martinsen, 1990).
Upward-coarsening facies successions can be produced by the High rates of deposition result in rapid burial and preservation
progradation of wave-formed shorefaces. However, unless of structures formed by unidirectional or oscillatory flows.
there is a significant supply of sediment, which almost invari- Sorting, especially in gravelly systems, may be poor to moder-
ably must come from rivers, significant progradation does not ate (Arnott, 1992; Bridge, 2003). Variations in discharge of the
occur. fluvial feeder system may produce an irregular coarsening-
upward succession, with interbedded mudstones throughout
River-Dominated Delta-Front and Prodelta Successions.— (Figs. 26, 28, 31).
Fresh-water influence may be indicated by an abundance of
In river-dominated deltas, prodelta sediments are typically syneresis cracks, reflecting flocculation and contraction of clays
heterolithic laminated to thin-bedded mudstones with or with- as a result of salinity changes (Plummer and Gostin, 1981), and
out sandstones (Fig. 14A, B). Siltstones and sandstones are early diagenetic siderite, which commonly requires fresh-water
typically massive to well stratified and may show graded bed- influx to reduce sulfate activity (Coleman and Prior, 1982;
ding (Fig. 28, river-dominated column). The graded beds may Bhattacharya and Walker, 1991b).
reflect deposition from hyperpycnal density underflows gener-
ated at the river mouth during high-discharge floods (Wright et Ichnological Effects of Fluvial Processes.—
al., 1988; Mulder and Syvitski, 1995). The degree of bioturbation
can be variable, depending on rates of sedimentation and the Periods of low discharge (e.g., hypopycnal) may result in
grain size of the sediment supplied (MacEachern et al., 2005; intense faunal colonization of the substrate alternating with
DELTAS 259

FIG. 25.—Percent-sandstone and net-sandstone thickness maps of the Eocene Slick Sand (a shelf-edge delta), Texas Gulf Coast,
U.S.A. Percent-sandstone map gives the best indication of the lobate nature of the delta. Growth faults (heavy lines on net-
sandstone-thickness map) have a fundamental control of facies distribution, with thickening on the hanging wall. After Winker
and Edwards (1983).

sparsely burrowed flood deposits, resulting in a highly irregu- feeders, such as are common in the Skolithos ichnofacies, and
lar, or even cyclic, bioturbation index (MacEachern et al., 2005). simple deposit feeders are more common. Buoyant plumes of
Preserved organic matter is commonly high in river-dominated sediment also create a sunlight stress, further reducing biogenic
delta fronts, reflecting numerous phytodetrital pulses of depo- reworking.
sition (MacEachern et al., 2005). Fresh-water or brackish-water
influence may be reflected in the trace faunal assemblages Wave-Influenced Delta Fronts.—
(Moslow and Pemberton, 1988; Bhattacharya and Walker,
1991b; Gingras et al., 1998; MacEachern et al., 2005). River- Wave-influenced deltas commonly consist of a series of pro-
dominated substrates produce the most stressful conditions for grading beach and shoreface complexes, with sand fed from a
infauna. Salinity stress leads to conditions that can be exploited nearby river (e.g., Rhone, Danube, Paraíba do Sul; Figs. 9, 17, 18).
only by trophic generalists. Ichnological suites are dominated Many wave-influenced deltas show an asymmetry that results
by low diversity and locally high abundance of generally dimin- from oblique wave approach, with amalgamated sandy beach-
ished forms. High amounts of suspended sediment inhibit filter ridge and shoreface deposits on the updrift side and muddier
260 JANOK P. BHATTACHARYA

FIG. 26.—Typical coarsening-upward facies successions formed


as a result of prograding deltaic lobes and mouth bars. Missis-
sippi example shows a composite of a thicker mouth-bar
succession below and the more irregular bay-fill successions
above. (From Bhattacharya and Walker, 1992; after Elliott,
1986, and Coleman and Wright, 1975). Rhone example (left)
shows progradation of a mouth bar forming an asymmetrical
coarsening-upward succession, whereas Rhone example
(right) shows truncation by a distributary channel, and has a
more symmetrical profile. Rhone examples are from
Bhattacharya and Walker (1992); modified by Elliott (1986)
after Oomkens (1970).

facies on the downdrift side (Figs. 6, 9). Updrift areas are usually
characterized by a relatively continuous coarsening-upward fa-
cies succession representing a wave-dominated shoreface (e.g.,
as in Fig. 28, wave-dominated column; Figs. 32, 33). The propor-
tion of wave-produced structures (such as wave ripples and
hummocky cross stratification) are greater updrift, whereas indi-
cators of high sedimentation rates and fresh-water influence (e.g.,
soft-sediment deformation, climbing current ripples, brackish
fauna) are fewer. Sandy sediment may be texturally more mature
and better sorted than on the downdrift side, where fluvial
influence is greater (Dominguez et al., 1987). Prodelta mudstones
may be more bioturbated, thinner, and sandier than in river-
dominated settings (Fig. 14D).
In the geological record, a single vertical facies succession of FIG. 27.—Examples of “forced” and “normal” regression (modi-
this type indicates a prograding wave-dominated shoreface. fied after Helland-Hansen and Gjelberg, 1994). A) Sharp-
Good three-dimensional control may be necessary before such a based shoreline deposits are produced when the trajectory of
shoreface can be positively ascribed to a delta. However, recent a falling shoreline is greater than sea-floor slope. B, C) Grada-
studies of many so-called “classic” shoreface successions are tionally based deposits, are predicted when falling shoreline
showing high ichnological stress in intervening mudstones, which trajectory is equal to or less than sea-floor slope. C)
may be a direct indicator of a brackish, fluvial plume nearby (e.g., Oversteepening can cause sediment gravity flows that are
Hampson and Howell, 2005). deposited on the basin floor. In all cases of forced regression
In asymmetric wave-influenced deltas, the sandy mouth bar (B, C, D) there is no subaerial accommodation and delta topset
may be elongated downdrift to form a barrier island, is seen in the facies are thin to absent. Thin topset facies may easily be
Brazos, Danube, and São Francisco deltas. The barrier may tem- reworked or eroded during subsequent transgression. D)
porarily inhibit delta progradation, because the back-barrier This contrasts with normal regression, where shoreline trajec-
lagoon area acts as an important area for trapping of river- tory is opposite of the basin slope. As a consequence, subaerial
derived sediment. Vertical facies successions may appear more accommodation is positive and significant accumulation of
like the irregular river-dominated examples described above. delta topset facies (i.e., fluvial channels, mudstones) can oc-
River-borne mud is deposited in greater proportions in the cur. Thick paralic and nonmarine facies thus accumulate and
downdrift than in the updrift areas (Fig. 32). are more likely to be preserved.
DELTAS 261

River Storm-Wave Storm-Wave


Dominated Influenced Dominated

Legend

FIG. 28.—Comparison of delta-front successions in river-domi-


nated, wave-influenced, and wave- dominated deltas in the
Upper Cretaceous Dunvegan Formation, Alberta, Canada.
After Bhattacharya and Walker (1991). The river-dominated
succession is the most irregular. Basal mudstones are in-
creasingly bioturbated with decreasing fluvial influence.

Tide-Influenced Delta Fronts.—


Tidally influenced delta fronts, such as the Fraser in Canada
(Monahan et al., 1997), the Mahakham in Indonesia (Allen et al.,
1979; Roberts and Sydow, 2003), the Niger in Africa (Allen, 1970),
the Fly delta in Papua New Guinea (Baker et al., 1995; Dalrymple
et al., 2003) the Mekong in Vietnam (Ta et al., 2002; 2005), and the
Baram and Trusan deltas in Borneo (Lambiase et al., 2003) also
show an overall coarsening-upward facies succession, but inter-
nally the facies reflect tidal influence. Tidal indicators in delta-
front sands include herringbone cross bedding, tidal bundles,
and reactivation surfaces, although these features are also found
in many nondeltaic tidal settings (Dalrymple, 1992).
Sandstones of the Frewens Allomember of the Frontier For-
mation in Wyoming, U.S.A., provide a recently studied example
of an ancient tide-influenced delta front (Figs. 34, 35; Willis et al.,
1999; Bhattacharya and Willis, 2001). Upward-coarsening facies
successions, 30 m thick, show an extremely low degree of burrow-
ing (Fig. 34) but contain marine dinoflagellates, indicating a
brackish-marine setting. Tidal features abound, including
heterolithic wavy-bedded mudstones and rippled sandstones at
the base, passing into thicker cross-bedded sandstones with
262 JANOK P. BHATTACHARYA

FIG. 29.—Deformation structures (load casts) in: A) prodelta mudstones of the Kavik Formation, Prudhoe Bay Field, Alaska, U.S.A.;
B) deformed sandstone bed overlain by parallel-laminated to rippled delta front splays interpreted as distal delta front, sediment-
gravity-flow deposits, Cretaceous Ferron sandstone, Utah, U.S.A.

abundant double mud drapes and reactivation surfaces (Fig. 35). numerous erosional features interpreted to have been produced
Mudstones show abundant subaqueous shrinkage cracks, which by tidal scours (Willis and Gabel, 2001). The tops of tide-influ-
may reflect salinity changes. However, these “syneresis” cracks enced deltaic successions are commonly reworked by tidal pro-
are also associated with small-scale interstratal deformation, cesses, producing deep tidal scours that might be mistaken for
which may indicate diastasis rather than syneresis (Cowan and fluvial or distributary-channel erosion surfaces. The Mahakham
James, 1992). delta contains 12 terminal distributary channels but over 20
The upward coarsening and low bioturbation indicate deltaic distinct bars. Bars not fed by active channels are bounded by
progradation, but the tidal features throughout indicate signifi- landward-narrowing tidal channels that scour up to 30 m deep
cant tidal modulation. The top of the Frewens sandstone is (Fig. 17; Allen et al., 1979). Other ancient examples of tide-
characterized by meter-thick sets of angle-of-repose cross beds, influenced deltas have been presented by Mutti et al. (1985),
commonly floored by mud chips. Vertical cliffs expose seaward- Maguregui and Tyler (1991); Nummedal and Riley (1999), and
dipping clinoforms (see Fig. 42) interpreted as seaward-migrat- Jennette and Jones (1995).
ing, tidally influenced mouth bars (Willis et al., 1999). In the case
that only cores were available through this system, these features Delta-Plain Successions
would look like erosionally based distributary channels.
Paleocurrents in the Frewens are dominantly unidirectional and Distributary Channels.—
strongly ebb-dominated, also suggesting tidal modulation of
river flows. Distributary channels are erosionally based (Fig. 36). Filling
Detailed work on the mixed tide- and wave-influenced Creta- commonly takes place after channel switching and lobe abandon-
ceous Sego sandstone in the Book Cliffs of Utah, U.S.A., shows ment. At this time, the distributary channel may develop into an
DELTAS 263

FIG. 30.—Cross section and interpretation of growth faults formed in prodelta and delta-front strata of the Cretaceous Ferron
Sandstone exposed along Muddy Creek, Utah, U.S.A. A) Detailed photomosaic, B) geological interpretation of structure, C)
detailed measured sections, and D) a reference diagram. The growth interval consists of upstream- and downstream-accreting
cross-bedded sandstones deposited in shallow distributary channels and proximal distributary-mouth bars. Successive sand-
stones in the growth section are labeled SS1 to SS6 (from Bhattacharya and Davies, 2001, 2004).
264 JANOK P. BHATTACHARYA

FIG. 31.—Cores and well logs from the modern Mississippi delta show a variety of upward-coarsening facies successions. Sharper-
based, blockier-appearing log patterns (e.g., profiles 5 and 8) lie in the more proximal portions of the delta lobe. Interlobe
succession (profile 1) is irregular. From Coleman and Prior (1982).

FIG. 32.—Block diagram illustrating the hypothesized three-dimensional facies architecture of an asymmetric delta. Significant
prodelta mudstones are associated with downdrift portion of the delta where sandy barrier-bar complexes occur within lagoonal
mudstones and bayhead-delta deposits. The updrift side of the delta comprises a sandy beach-ridge plain (from Bhattacharya and
Giosan, 2003).
DELTAS 265

FIG. 33.—Facies photos of wave-dominated shoreface of the


Gallup Sandstone, New Mexico, U.S.A. A) Distance shot
of wave-dominated shorefaces shows sand-dominated
cliff section. B) Close-up of basal sandstones showing
pervasive bioturbation. Large mud-rimmed burrow is
Asterosoma. Smaller sand blebs in mud rim are Chondrites.
C) Mud-pellet-lined Ophiomorpha burrows in cross-bed-
ded shoreface sandstones in middle part of cliff. These
suggest a wave-dominated shoreface characterized by the
Skolithos ichnofacies. D) Bidirectional cross bedding in the
upper shoreface.

estuary, and the fill is commonly transgressive with strong tidal depends on the degree of river dominance and the position of the
indications. The facies succession tends to fine upward, with channel. Distributary channels within the nonmarine upper delta
some preserved river-derived facies at the base and a greater plain look entirely fluvial in nature (see Bridge, 2003, and Bridge,
proportion of marine or brackish facies in the upper part of the this volume), although they tend to be single-story rather than
channel fill. The extent of brackish to marine facies development multi-story, compared to valley fills.
266 JANOK P. BHATTACHARYA

FIG. 34.—Measured vertical facies succession through tide-influenced delta front of the
Cretaceous Frewens Allomember, Wyoming, U.S.A., emphasizes low degree of
burrowing and heterolithic nature of interbedding. Photos of facies in Figure 35.
From Willis et al. (1999).

Although the salt-water wedge migrates no farther landward


than the bay line, tidal effects can be felt farther upstream. As a
consequence, rhythmic alternations of mud and sand may be
seen in wholly freshwater channels that nevertheless indirectly
feel some marine influence (Gastaldo et al., 1995). Examples of
these different types were presented by Bhattacharya (1989) and
Bhattacharya and Walker (1991b) from distributaries in Creta-
ceous deltaic systems of the Dunvegan Formation in Alberta,
Canada (Fig. 36).
The overall proportion of distributary-channel facies is a
function of the type of delta, avulsion frequency, bifurcation
order, and channel migration (e.g., Bristow and Best, 1993; Miall,
1996; Blum and Törnqvist, 2000; Bridge, 2003; Bridge, this vol-
ume; Olariu and Bhattacharya, 2006). Numerical models of allu-
vial versus deltaic stratigraphic systems do not typically allow
multiple active distributaries, but rather model a single channel
DELTAS 267

FIG. 35.—Tide-dominated delta front of the Cretaceous Frewens sandstone, Wyoming, U.S.A. A) Upward-coarsening facies
succession (see measured section in Figure 34). B) Double mud drapes indicative of tidal modulation. C) Heterolithic, lightly
burrowed subtidal prodelta facies at the base of the succession. Bedding architecture is shown in Figure 42.

that avulses and migrates (e.g., Paola, 2000; Mackey and Bridge, finger with adjacent floodplain or delta-plain facies but exhibit
1995; Overeem and Weltje, 2001; Overeem et al., 2005; Olariu and an erosional relationship. This may be more difficult to observe
Bhattacharya, 2006). Some of the fundamental characteristics of in well-log data, and large valleys may internally contain flood-
distributive channel systems, particularly downstream decreases plain or delta-plain facies. Interfluve paleosols may also pro-
in discharge, channel width, and channel depth, are thus not vide key evidence of sediment bypass, floodplain starvation,
predicted by present numerical models, and the increase in and avulsion frequency (McCarthy, 1999; Kraus, 2002).
bedload-related form friction is not generally accounted for
(Giosan and Bhattacharya, 2005; Overeem et al., 2005). Interdistributary Areas.—
In general, the more wave-dominated the delta, the greater is
the proportion of lobe sediment, with more limited amounts of Interdistributary and interlobe areas are less sandy, and com-
interlobe and distributary-channel facies. It is also emphasized monly contain a series of relatively thin, stacked coarsening- and
that there is no such thing as “a distributary channel” of a unique fining-upward facies successions (Figs. 31, 39). These are usually
width or depth. Several scales of distributary channel may occur less than ten meters thick, and they do not show as thick or as
within any given delta (Fig. 18). well-developed coarsening-upwards facies successions as are
In general, valley fills are much thicker than associated found in prograding deltaic lobes (Elliott, 1974; Tye and Coleman,
delta-front successions and consist of multi-story sandstones 1989). The proportion of lobe versus interlobe successions de-
(Figs. 19, 37; Reynolds, 1999). Plan-view maps of valleys should pends on the nature and type of delta system, the stability of
show a tributive rather than a distributive pattern (Fig. 38; e.g., distributary channels, and the amount of nonmarine accommoda-
Plint and Wadsworth, 2003). While these may be difficult to map tion. Wave-influenced systems, like the Danube, can contain sig-
in outcrops or in sparse subsurface data sets using 1D log or 2D nificant lagoonal and bay mudstones in regions downdrift of the
seismic data, horizon mapping through 3D seismic cubes has river mouth, and, depending on shoreline trajectory, thick accu-
revealed complex dendritic drainage patterns within buried mulations of mud-prone paralic and nonmarine facies can accu-
paleovalleys (e.g., Brown, 2005). Also, valley fills do not inter- mulate behind an aggrading shoreface or delta front (Fig. 27D).
268 JANOK P. BHATTACHARYA

Interdistributary Areas in River-Dominated Deltas.—


An interdistributary bay is filled by overbank spilling of fine-
grained material from the river during flood stages (Fig. 8). There
is an overall shallowing-upward facies succession, associated
with a trend from more marine to more nonmarine facies, but
commonly without the deposition of thick sands (Fig. 39). This
represents the transition from offshore prodelta mudstones into
delta-top facies without the development of a sandy shoreline
(Walker and Harms, 1971; Bhattacharya and Walker, 1991b). The
muddy nature of interdistributary-bay successions may be punc-
tuated by sandy crevasse-splay or channel deposits that may
produce thin coarsening or fining successions (Coleman and
Prior, 1982; Elliott, 1974). The succession may grade into rooted
coaly mudstones or coals representing a variety of swamp, marsh,
and lacustrine environments. Beach sands, associated with the
development of barrier strandplains, spits, or cheniers, may be
present at the tops of these successions, although they are rela-
tively thin compared to the delta front. Interlobe areas may also
act as the locus for progradation of a subsequent lobe and may be
erosionally truncated by younger distributary channels.


FIG. 36.—Comparison of distributary-channel-fill successions in
river- and marine-dominated deltas of the Dunvegan Forma-
tion (Cretaceous, Alberta, Canada). In the marine-dominated
system, the distributary fill reflects transformation of the
distributary into an estuary. After Bhattacharya and Walker
(1991b). Legend in Figure 28.

C
C
B
B
A A

FIG. 37.—Facies architecture of an interpreted valley fill in the Cretaceous Ferron Sandstone member of the Mancos Shale, Utah,
U.S.A. Base of valley erodes into several upward-coarsening parasequences A, B, C) Valley depth (Hv) is about 21 meters. In
contrast, associated channel depths (Hc) are only about 6 m. Valley is filled with 5 channel stories (1–5). Lowest channel-belt
deposit (1) is largely eroded by migration of younger channels. Predominance of laterally accreting bars defines the internal
facies architecture of each channel-belt deposit. The bedding geometry shows that the rivers were single-thread, meandering
streams that gradually filled the larger valley. Calculation of water depth from dune-scale cross strata within the bar deposits
suggest maximum bankfull depth of about 9 m. From Bhattacharya and Tye (2004), modified after Barton et al. (2004).
DELTAS 269

A
N
200 km

Perm

Moscow

Tributive
TributiveSystem
System
(Volga
(Volga
V Drainage
DrainageBasin,
3Basin,
2
Area
Area1,614.4 x 103 km
1,614.4x10 km2))

Area dominated by
erosion processes
"Trunk" Channel (Volga River),
might be extremely short for some systems
Area with erosion Volgograd
or deposition
Apex
Distributive System
(Volga Delta,
Area dominated by 2
Astrakhan Subaerial area 27,224 km )
deposition processes

Basin
(Caspian Sea)

gan Delta)
ed

FIG. 38.—A) Example of a tributive–distributive system, Volga basin. The tributive pattern is an order of magnitude larger (tens to
hundreds of times) than the distributive pattern, and the main “trunk” valley connects the two patterns. Modified after Payne
et al. (1975). B) Tributive-trunk system in Dunvegan lacks details of distributive pattern because distributary channels are too
small to image. (From Plint and Wadsworth, 2003).
270 JANOK P. BHATTACHARYA

Interdistributary Areas in Tide-Influenced Deltas.—


Tidal processes may be important in interdistributary bays
(even in river-dominated deltas), resulting in tidally influenced
facies such as tidal flats or tidal channels (Allen et al., 1979; Ramos
and Galloway, 1990). These are especially common in modern
tidally influenced deltas such as the Niger, Fraser, Mahakham,
Fly, Mekong, Ganges–Brahmaputra, Ayerarwady, and Orinoco
deltas. Ancient examples of tidally influenced facies in delta-
plain settings include those published by Ramos and Galloway
(1990), Eriksson (1979), and Rahmani (1988). Ebb versus flood
tidal currents may move down different pathways, such that
current directions are unidirectional at any one place but differ
between ebb-dominated versus flood-dominated channels (Har-
ris, 1988; Dalrymple et al., 2003). In tide-dominated deltas, mud is
partitioned along the sides of the system as well as in the prodelta
area (Fig. 20; Dalrymple, 1992; Willis et al., 1999; Willis, 2005).

FACIES ARCHITECTURE OF DELTAS


Bedding geometry and lateral facies variability can be ad-
dressed by the use of seismic data (e.g., Hart and Long, 1996;
Anderson, J.B. et al., 2004), ground-penetrating radar (Jol and
Smith, 1991, 1992; Smith et al., 2005; Lee et al., 2005), continuous
outcrop data (e.g., Willis et al., 1999; Soria et al., 2003; Gani and
Bhattacharya, 2005; numerous papers in Chidsey et al., 2004),
and interpolation of well data (e.g., Bhattacharya, 1991, 1993,
1994; Ainsworth et al., 1999; Tye et al., 1999; Plint, 2000;
Bhattacharya and Willis, 2001). Facies architectural studies of
deltas lag significantly behind those of fluvial, deep-water, and
eolian systems, but more recent studies of deltaic systems are
becoming available (Willis et al., 1999; Knox and Barton, 1999;
Willis and Gabel, 2001; Olariu et al., 2005).
Although it may be premature to fully characterize the archi-
tectural elements that make up deltaic depositional systems,
some generalizations can be made. Sandy architectural elements
in the delta plain include channels at various scales, which may
migrate or stack to form channel bodies or channel belts. Inter-
nally channel bodies consist of bars (macroforms) and smaller-
scale bedforms, analogous to the architectural elements described
in the fluvial literature (Miall, 1995, 1997; Bridge, 2003; Bridge,
this volume). The number of different scales of channels relates to
FIG. 39.—Interdistributary-bay fill in a river-dominated delta lobe the bifurcation order, which can be high in river-dominated
in the Dunvegan Formation (Cretaceous, Alberta, Canada), deltas and low in wave-dominated settings. Unfortunately, bifur-
showing thin irregular cycles and overall increase in propor- cation order is very difficult to determine in outcrop or subsur-
tion of nonmarine facies upwards. Compare with profile 1 in face examples, although this may be possible in selected settings
Figure 31. Legend in Fig. 28. After Bhattacharya and Walker or particularly good outcrop exposures (e.g., Bhattacharya and
(1991). Tye, 2004). Areas away from distributary channels may include
crevasse splays and levee deposits. The delta plain also consists
of numerous mud-prone wetland environments, although there
Interdistributary Areas in Wave-Influenced Deltas.— have been few studies that compile the typical dimensions of the
associated muddy facies elements.
Interdistributary bays may often be closed off by barrier– The distal delta plain and proximal delta front consist of
beach complexes in wave-dominated deltas resulting in exten- mouth-bar elements, which in turn build bar assemblages and
sive back-barrier lagoons — e.g., the Nile (Fig. 1), the São Fran- form depositional lobes. There may be several scales of bar
cisco, the Danube (Fig. 9), the Brazos, and the Po. These may be assemblage and lobe clustering, especially in continental-scale
filled from the landward side by progradation of bayhead deltas river-dominated delta systems like the Mississippi. Mouth bars
or from the barrier side by storm washovers (Bhattacharya and are in turn intimately associated with terminal distributary chan-
Giosan, 2003) and have been regarded as local estuaries within a nels. A variety of sandy bedforms may be associated with the
larger delta system. Deposits are commonly organic-rich, with upstream sides of these bars. In river-dominated, shallow-water,
marsh vegetation or mangroves. These areas are typically more friction-dominated deltas, these channels are typically only a few
pronounced on the downdrift sides of highly asymmetric deltas meters in depth and a few tens to a few hundred meters wide. The
(Bhattacharya and Giosan, 2003). However, identification of es- distal margins of bars are commonly formed by frontal splay
tuarine-type facies does not necessarily mean deposition within elements or subaqueous channels and chutes. Channels typically
a valley (e.g., MacEachern et al., 1998). scour only a few meters and may be intimately associated with
DELTAS 271

frontal splays. Dimensions of splays are largely unknown, al- for hundreds of kilometers along depositional strike? The posi-
though individual beds should scale to the generative flow. These tion of river plumes in these outcrops is indicated by the appear-
frontal splays may coalesce to form a fringe of distal delta-front ance of steep clinoform strata, a marked decrease in diversity and
sand, which scales to the size of the depositional lobe. abundance of ichnofauna in underlying mudstones, and the
Wave-formed architectural elements include barrier-island appearance of laminated to graded prodelta facies (Hampson
sand bodies and shorefaces. Width of barrier islands typically and Howell, 2005).
scales to the width of the initial mouth bar, although they may Tidal reworking may produce a bewildering variety of tidal
extend for several kilometers downdrift. Shorefaces can reach bedforms and bars (see Dalrymple, 1992; Boyd et al., this vol-
several to tens of kilometers in width and several tens of kilome- ume; Willis, 2005). Tidal processes also result in greater win-
ters in length. In many wave-dominated coastlines, the area nowing of distributary channels, which may be stable for con-
occupied by shoreface “wings” can greatly exceed the area occu- siderably longer periods than terminal distributary channels in
pied by the river-dominated mouth bar (e.g., Paraíba do Sul, Fig. river-dominated environments. This can result in significantly
18B). This has led many to question the value of calling river-fed more elongate bars and bar assemblages than in nontidal set-
shorefaces, deltas at all (Dominguez, 1996). Recent studies of the tings.
shoreface successions in the Book Cliffs, Utah, U.S.A., suggests Muddy elements associated with the subaqueous realm in-
that rivers were widely spaced, up to 50 kilometers apart, along clude prodelta muds, bay muds, and bar drapes. Prodelta muds
strike (Hampson and Howell, 2005). The shorefaces are charac- may cover vast areas of the shelf and may migrate for thousands
terized by river-plume deposits that effectively “puncture” the of kilometers along strike. Although muddy wave-formed clastic
otherwise rather uniform shoreface sand body. These studies coastlines are common in the modern, there is a paucity of well
predict that shoreface sandstones typically extend several tens of documented ancient examples such as cheniers. Tidal flats are
kilometers along strike before they are punctured. This invites the significantly better recognized but are largely discussed in the
question: are there any shorefaces that truly extend, unbroken, context of tidal depositional systems.

FIG. 40.—Inclined bedding (clinoforms) and facies in a river-dominated delta front of the Cretaceous Ferron sandstone member, Utah,
U.S.A. A) Photomosaic of a cliff face. B) Bedding and facies geometry of the same cliff face (along depositional dip), Ivie Creek
amphitheater, Emery County, Utah. The diagram shows prominent seaward-dipping clinoforms. From Gani and Bhattacharya
(2005), modified after Anderson, P.B. et al. (2004) and Mattson (1997).
272 JANOK P. BHATTACHARYA

A comprehensive compilation of the sizes and dimensions of Deltaic deposits are characterized by a prograding clinoform
these various architectural elements is simply beyond the scope geometry (Figs. 40–46; (Gilbert, 1885; Barrell, 1912; Rich, 1951;
of this review, and in this regard architectural-element analysis of Scruton, 1960; Berg, 1982). This geometry can be seen in downdip
deltaic systems remains significantly behind the deep-water, seismic profiles of modern and ancient deltas (see Figures 43 and
fluvial, and eolian systems described in other chapters in this 44). Many superb examples are given in a recent volume on
volume. Many of the elements described in these other systems Quaternary deltas of the Gulf of Mexico (Anderson and Fillon,
are also found in deltas (e.g., channels, bars, shorefaces, barriers, 2004), as well as numerous examples from Southeast Asian deltas
tidal bars, cheniers), but clearly a great deal of compilation is presented in Sidi et al. (2003). This clinoform geometry can also be
required to determine whether the dimensions of these elements reconstructed in core and well-log cross sections (Figs. 45, 46;
is fundamentally different if they are associated with a delta Bhattacharya, 1991; Ainsworth et al., 1999; Plint, 2000). It can also
system. Tye (2004) compiled data on the dimensions of mouth- be seen in some outcrops (Figs. 40–42; Chidsey et al. (2004); Gani
bar sandstones in several arctic deltas, as well as the Atchafalaya and Bhattacharya, 2005). Berg (1982) discussed typical seismic
delta in the Gulf of Mexico, and Reynolds (1999) compiled data on facies in deltaic depositional systems and suggested that sandy
dimensions of a variety of paralic sandstone bodies, including wave-dominated systems are characterized by a shingled pat-
mouth bars and distributary channels. tern, whereas muddier deltas show an oblique-sigmoidal pat-
tern. Sigmoid-shaped portions are characteristic of the mud-
Dip Variability dominated prodelta facies (Kuehl et al., 1997; Liu et al., 2002;
Roberts and Sydow, 2003; Hiscott, 2003; Anderson, J.B. et al.,
In cross-sectional dip view, deltas can be divided into three 2004; Roberts et al., 2004; Neill and Allison, 2005; Kuehl et al.,
distinct regions: topset, foreset, and bottomset (Figs. 2, 3, 40). 2005), whereas the more flat-lying or oblique reflectors represent
Foresets are typically associated with the distal delta front and the sandier delta-front and delta-plain facies. Frazier (1974)
show dips that can range from a few degrees up to the angle of showed a similar clinoform geometry on the basis of geological
repose in Gilbert-type deltas (Fig. 41). Bottomsets dip less steeply studies of the Mississippi delta plain.
than foresets, typically << 1°, and are usually associated with Offlapping clinoformal geometries have also been recognized
prodelta sediments. Topset facies are typically flat to undulating in Late Quaternary deltas around the world (Brown and Fisher,
and are built out of the proximal delta-front and delta-plain 1977; Suter and Berryhill, 1985; Tesson et al., 1993; Sydow and
facies. Subaqueous topset facies (i.e., proximal delta front) can Roberts, 1994; Hart and Long, 1996; Hiscott, 2003; Roberts and
show extreme variability, depending on whether they are con- Sydow, 2003; Roberts et al., 2004; Bart and Anderson, 2004; and
structed from mouth bars, tidal bars, or shoreface deposits. others in Anderson and Fillon, 2004). In shelf-edge systems,
Shoreface elements produce the simplest bedding geometries clinoforms commonly steepen towards the shelf edge. The steep-
(e.g., Hampson, 2000) and essentially consist of seaward-dipping ening reflects the progradation into progressively deeper water.
beds, with dip angles typically less than 1°. Mouth bars and tidal At the shelf edge, the delta front can no longer build seaward, so
bars produce far more complex deposits (Fig. 42). The seaward it steepens and then fails (e.g., Fig. 24).
migration of these elements builds the vertical facies successions Clinoform gradients have a wide range of values in different
detailed above Fig. 34). settings. Clinoform gradients of shelf-edge deltas in the Gulf of

FIG. 41.—Details of facies interfingering at the base of a small-scale outcrop example of a gravelly, Pennsylvanian “Gilbert” delta, Taos
Trough, New Mexico, U.S.A. A) Outcrop photomosaic. B) Line drawings of beddings with facies interpretation. Note that
clinoforms are steeply dipping (average 13°). C) Lithologic column of this coarse-grained delta (position of the measured section
is shown in Part B). From Gani and Bhattacharya (2005).
DELTAS 273

FIG. 42.—Outcrop example of complex internal architecture in the Cenomanian (Upper Cretaceous) tide-influenced river delta of the
Frewens Allomember, Frontier Formation, central Wyoming, U.S.A. Dip view (AB) of the prograding delta shows the seaward-
dipping clinoforms, whereas in strike view (BC) these clinoforms show bidirectional downlap, forming a classical lens-shaped
geometry. In both cases, muddy bottomset facies interfinger with the sandy foreset facies, forming a shazam-type facies
boundary. Note that clinoform dip varies from 5° to 15°. Detailed facies shots and measured sections are shown in Figures 34 and
35 (modified from Willis et al., 1999).
274 JANOK P. BHATTACHARYA

FIG. 43.—A) Dip-oriented and B) strike-oriented views showing bedding geometry of a top-truncated lowstand delta, based on
shallow seismic profiles off the Natashquan River, Gulf of St. Lawrence, Canada (after Hart and Long, 1996). Note reworked
sediments on top of deltas.

Mexico average between 4° and 8° (Suter and Berryhill, 1985). tectonics, because deltas commonly fill low areas on the sea floor
Gradients of the Rhone shelf-edge deltas average about 1°. The (Bhattacharya and Willis, 2001; Martinsen, 2003). Tectonics can be
high slopes in the Gulf Coast and Rhone shelves result in signifi- related to salt or shale mobility, or it can be related to larger-scale
cant instability of the shelf-edge sediments, where large-scale lithospheric deformation (e.g., plate tectonic).
synsedimentary deformation features are common. Slopes are
typically much lower in ramp-type depositional margins, and Sequence Stratigraphy of Deltas
soft-sediment deformation features in the Alberta examples are
mostly limited to loading rather than large-scale slumps, slides, Delta systems have been an important focus of research in the
or growth faults. Foreset dips of subaqueous prodelta clinoforms development of new allostratigraphic and sequence stratigraphic
are typically less than 0.1° (e.g., Liu et al., 2002; Neill and Allison, concepts (e.g., Boyd et al., 1989; Van Wagoner et al., 1990;
2005) as opposed to sandy foresets, which are usually an order of Posamentier et al., 1992; Bhattacharya, 1993; Miall, 1997;
magnitude higher (i.e., > 1°). Posamentier and Allen, 1999; Anderson, J.B. et al., 2004). Se-
quence stratigraphy provides a very different view of both the
Strike Variability large-scale and small-scale architecture of sedimentary systems
(e.g., Van Wagoner et al., 1990; Bhattacharya and Posamentier,
Along strike, facies relationships may be less predictable and 1994; Miall, 1997).
depositional surfaces may dip in different directions (Figs. 42, Allostratigraphy and sequence stratigraphy involve the corre-
43B, 44B, 47). This is particularly so in more river-dominated lation of bounding discontinuities through potentially varying
deltas, where along-strike reworking is not significant and abrupt lithologies that yields a fundamentally different picture of genetic
facies transitions may occur between distributaries and stratal relationships than older lithostratigraphic techniques, and
interdistributary areas (Bhattacharya, 1991). Overlapping delta allows far more accurate paleogeographic maps to be constructed
lobes result in lens-shaped stratigraphic units that exhibit a (Fig. 48). This is best illustrated by example. Historically the term
mounded appearance on seismic lines (Figs. 43B, 44B). Regional delta has been generally applied to many clastic wedges, such as
mapping of delta lobes in the Cretaceous Dunvegan Formation, the Devonian–Carboniferous Catskill delta wedge in the north-
Alberta, Canada (Bhattacharya, 1991; Plint, 2000), the Frontier eastern U.S.A. (Woodrow and Sevon, 1985) and the Cretaceous
Formation in Wyoming, U.S.A. (Bhattacharya and Willis, 2001), Dunvegan, Ferron, and Frontier formations in Western North
and Pennsylvanian deltas in Kentucky (Horne et al., 1978) shows America. Previous lithostratigraphic maps of these undifferenti-
similar lateral overlap of lens-shaped delta lobes (e.g., Fig. 47). ated wedges show broadly lobate geometries, especially at the
Although the older seismic and sequence stratigraphic litera- distal margins of the wedges, but it was practically impossible to
ture is rife with dip-oriented cross-sectional depictions of shelf determine which lobe belongs to which channel without more
depositional systems, newer studies emphasize strike-oriented detailed sequence stratigraphic correlations (Fig. 49).
variability (e.g., Anderson, J.B. et al., 2004). Strike variability (i.e.,
the timing and spacing of overlapping lenses or lobes) is depen- The Dunvegan Delta
dent on the number, spacing, and avulsion frequency of distribu-
tary channels. It also depends on the shape of the sea floor, The Dunvegan Formation of Alberta, Canada, represents a
especially if there is differential subsidence or uplift related to heterolithic wedge of mudstones and sandstones, up to 300 m
DELTAS 275

FIG. 44.—Seismic geometry of the Lagniappe delta, Gulf of Mexico. A) Dip line showing clinoforms; B) strike line showing lens-shaped
cross sections; C) base map showing outline of lobe; D) detailed seismic facies mapping shows sub-lobes and distributary
channels. Core MP303 c1 shows a predominantly upward-coarsening facies succession. After Roberts et al. (2004).
276 JANOK P. BHATTACHARYA

FIG. 45.—A) Dip-oriented well-log and core cross sections within Allomember E of the Upper Cretaceous Dunvegan Formation,
Alberta, Canada, showing offlapping clinoforms. Paleogeographic maps of the various offlapping shingled units are shown in
Figure 51. Modified after Bhattacharya (1991).

FIG. 46.—Subsurface model of well-log correlation in a lacustrine deltaic environment along depositional dip. Lithostratigraphic
correlation (upper diagram) assumes no dip in sand bodies towards basin, whereas chronostratigraphic correlation (lower
diagram) assumes basinward-dipping clinoforms. Chronostratigraphic model better predicts reservoir behavior. Note that
correlation lengths of many beds are below the well spacing (from Gani and Bhattacharya, 2005, modified after Ainsworth et
al., 1999).
DELTAS 277

FIG. 47.—A) Maps and B) cross section of delta lobes and fingers in the Cretaceous Lower Belle Fourche Member, Frontier
Formation, Wyoming, U.S.A. Strike section (b) shows overlapping lens-shaped delta bodies. Facies details of the Frewens tide-
dominated delta sandstone are shown in Figures 34, 35, and 42. (After Willis et al., 1999, and Bhattacharya and Willis, 2001.)

thick, deposited from the actively rising Western Cordillera into systems, including some superb examples of ancient river-domi-
the adjacent Cretaceous Interior Seaway. The term “Dunvegan nated deltas (Bhattacharya, 1991). The abundance of core data
Delta” has been applied to this entire undifferentiated sedimen- allowed reconstruction of the lateral facies relationships both
tary package (Figs. 48, 49). down dip (Fig. 45) and along depositional strike. The cores also
In a study area of about 30,000 km2 in the subsurface of facilitated the development of summary vertical facies succes-
Alberta, Bhattacharya and Walker (1991) recognized seven sions for the various components in the different deltaic systems
throughgoing transgressive surfaces (Fig. 48B). These were used (Figs. 28, 36, 39).
to subdivide the Dunvegan wedge into seven allomembers (Fig. Individual delta lobes could not be mapped without the
48B). Each of the allomembers could be further subdivided into detailed correlation of the offlapping, shingled units (Figs. 45, 50,
several shingled offlapping units separated by less extensive 51). If sandstones within each allomember are mapped together,
surfaces of transgression and regression (Figs. 45, 48). The discon- the isolith patterns do not show narrowing and thickening towards
tinuity-bounded shingles and allomembers provided the strati- a point fluvial source in a landward position (Fig. 49). Only the
graphic basis for more detailed facies mapping and paleogeo- seaward deltaic promontories can be seen (e.g., Fig. 49). Also, the
graphic reconstruction than had ever before been possible (Figs. highstand and lowstand deltas within each allomember could not
50–52 ). Continued work in a more landward direction by Plint be mapped without the detailed correlation of the shingle bound-
(2002) has demonstrated older shingles and allomembers (Fig. aries. This is an especially acute problem in high-accommodation
53) below those mapped by Bhattacharya and Walker (1991a). settings, where lowstands tend to be attached to the highstand
Sandbody geometries within individual shingles (Figs. 50, 51) clastic wedge, versus low-accommodation settings, where low-
revealed a wide range of deltaic to shoreline-related depositional stands are highly detached (Ainsworth and Pattison, 1994).
278 JANOK P. BHATTACHARYA

A Lithostratigraphy
Kaskapau Formation

Dunvegan
Formation La Biche

Shaftesbury Formation

Fish Scales Marker bed

B Allostratigraphy
Kaskapau Formation
Dunvegan Fm.

Shingle
Shaftesbury

50m
50 m

SANDSTONE 20km
20 km

FIG. 48.—Regional cross section across the Alberta Foreland Basin, Canada, illustrating the difference between a lithostratigraphic
and an allostratigraphic interpretation of the Upper Cretaceous Dunvegan Formation (from Bhattacharya, 1993). A) The
lithostratigraphic interpretation depicts a homogeneous, wedge-shaped sandstone body that tapers to the right. Some
interfingering of the distal end of the Dunvegan Formation into the La Biche Formation shales is shown. B) The allostratigraphic
interpretation shows that the Dunvegan comprises several stacked allomembers (A to G). Each allomember consists of several
smaller-scale, offlapping, shingled units that map as delta lobes (e.g., Figs. 50 and 51). Each allomember is bounded by a
regional transgressive flooding surface. These regional flooding surfaces and smaller-scale “shingle” boundaries show that the
Dunvegan consists of numerous sandy compartments, bounded by mudstones. Oil and gas reservoirs occur within these
smaller-scale shingled units.

More recent work (Plint and Wadsworth, 2003) in the land- surfaces, erosional surfaces, coal beds, bentonites, and ammo-
ward realm has allowed mapping of the nonmarine facies within nite horizons across the outcrop belt. The Ferron has been
the Dunvegan, including superb examples of tributive valley subdivided into seven major transgressive–regressive “strati-
systems (Fig. 38B). graphic cycles”, each of which is bounded by a regionally
traceable flooding surface and associated coal (Ryer, 1984;
Ferron Example Gardner, 1995). The lower two stratigraphic cycles consist of
strongly seaward-stepping shoreline and delta deposits. The
A recent correlation of the Ferron sandstone member, in middle three cycles aggrade, and the last two cycles backstep.
central Utah, U.S.A., based on nearly continuous outcrop expo- Regionally, the Ferron delta prograded to the northeast, but
sures, shows a complicated series of seaward-stepping, locally, individual delta lobes prograded at high angles to this
offlapping, to aggrading and finally backstepping shorelines general northeast direction.
(Fig. 54; Gardner, 1995; Gardner et al., 2004; Barton et al, 2004; Internally, the stratigraphic cycles consist of a series of lens-
Garrison and van den Bergh, 2004). The correlation is based on shaped to lobate offlapping, shingled delta-front and shoreface
tracing various bounding discontinuities, including flooding sandstone bodies that show upward-coarsening facies succes-
DELTAS 279

of the fault zones shows that deformation was largely by soft-


sediment mechanisms, such as grain rolling and by lubrication of
liquefied muds, causing shale smears. Mechanical attenuation of
thin beds occurs by displacement across multiple closely spaced
small throw faults. Analogous river-dominated deltaic subsur-
face reservoirs may be compartmentalized by growth faults, even
in shallow-water, intracratonic, or shelf-perched highstand del-
tas. Reservoir compartmentalization would occur where thicker
homogeneous-growth sandstones are placed against the muddy
pre-growth strata and where faults are shale-smeared, and thus
potentially sealing.
Although the Dunvegan and Ferron were deposited into a
very similar tectonic and paleogeographic setting, the Ferron
shows considerably more complexity in cross section (compare
Figures 53 and 54) . This likely reflects the fact that the Ferron is
nearly continuously exposed in outcrop, allowing greater reso-
lution of individual delta lobes than is possible with the largely
well-log-based cross section of the Dunvegan. Outcrop ex-
amples, like the Ferron, indicate the fundamentally complex
nature of reservoir compartmentalization within fluviodeltaic
reservoirs.

Top-Truncated Deltas and the Shelf Sand Problem

FIG. 49.—Sandstone isolith mapping of undifferentiated Dunvegan One of the central debates in facies sedimentology in the past
wedge (after Burk, 1963). few decades has been the interpretation of seemingly isolated
basin-distal, “shelf” sandstones (cf. Snedden and Bergman,
1999). These deposits comprise meters-thick to tens-of-meters-
sions. Wave-influenced and river-dominated successions can be thick, upward-coarsening facies successions that are surrounded
recognized. The shingled units are locally bounded by less exten- by “shelf” mudstone. Coevel paralic facies, such as typify the
sive flooding surfaces, and they thus resemble “parasequences” high-accommodation and unequivocally deltaic Dunvegan and
(e.g., Van Wagoner et al., 1990). The “stratigraphic cycles” effec- Ferrron sandstones described above, are commonly found tens
tively are “parasequence sets”. The shoreline-type sandstone to hundreds of kilometers away. In the 1970s and 1980s these
bodies correlate landward into a thick delta-plain succession, basinally isolated sandstones were widely interpreted as off-
consisting of single-story distributary channels and multi-story shore shelf bars, molded by shelf processes unrelated to the
valley fills (Fig. 37; Garrison and van den Bergh, 2004; Barton, shoreline, or deposited as turbidites, implying deeper-water
2004). Highly asymmetric subsidence, characteristic of the fore- depositional systems (see summary in Bhattacharya and Willis,
land-basin setting, results in high accommodation, high sedi- 2001).
ment supply, a broadly positive shoreline trajectory, and conse- In the late 1980s and the 1990s, sequence stratigraphic con-
quently accumulation and preservation of thick paralic, delta- cepts forced a reconsideration of these units, and it was recog-
plain topset facies (compare Fig. 54 and Fig. 27). nized that the sandstones, in many cases, were underlain by
The integration of detailed facies analysis within a well devel- erosional surfaces, implying no genetic link between the under-
oped stratigraphic framework allows specific statements to be lying shelf mudstones and the newly interpreted overlying
made about how depositional systems evolve as a function of shoreface deposits, interpreted to be deposited by the process of
position within the wedge. The seaward-stepping deltaic para- “forced regression” during major drops of sea level (Plint, 1988;
sequences in the lowest stratigraphic cycle show greater river Posamentier et al. 1992).
influence than the more wave- and tide-influenced parasequen- The next level of debate then centered on the magnitude of
ces in the upper cycles, although there is considerable facies sea-level drops and the origin of the erosional surfaces. Several
variability at all scales. studies went the ultimate step and interpreted the erosion sur-
The lower parasequences in the Ferron contain normal growth faces to be fluvially eroded, incised-valley fills (e.g., Jennette and
faults well exposed along the highly accessible walls of Muddy Jones, 1995). Thus, sediment bodies once interpreted as deeper-
Creek Canyon in Central Utah (Fig. 30; Bhattacharya and Davies, water “shelf” turbidites became fluvially incised valley fills.
2001, 2004). Distinctive pre-growth, growth, and post-growth An unwillingness to interpret these systems as strictly deltaic
strata indicate a highly river-dominated crevasse delta that pro- can be traced directly to the application of Barrell’s 1912 criterion,
graded northwest into a large embayment of the Ferron shore- namely that subaerial facies are required to define a delta. Strict
line. The growth section comprises medium- to large-scale cross- adherence to this rule does not consider the fact that during
stratified sandstones deposited as upstream- and downstream- forced regression, shoreline trajectory is negative, and accommo-
accreting mouth bars in the proximal delta front. Deposition of dation for delta-plain facies is extremely limited (Fig. 27). Also,
mouth-bar sands initiates faults. Because depositional loci shift ravinement is a very efficient process that truncates the tops of
rapidly, there is no systematic landward or seaward migration of these systems.
fault patterns. During later evolution of the delta, foundering of The Cretaceous Frontier Formation, in the Powder River
fault blocks creates an uneven sea-floor topography that is Basin of Wyoming, U.S.A., is a recently described example that
smoothed over by the last stage of deltaic progradation. Faults are contains numerous basin-distal sandstones separated by mud-
inferred to occur within less than 10 m water depths in soft, wet stones (Bhattacharya and Willis, 2001). Previous interpretations
sediment (Bhattacharya and Davies, 2004). Detailed examination of these sandstones, based on only limited data, ranged from
280 JANOK P. BHATTACHARYA

FIG. 50.—Paleogeographic maps of successive offlapping shingles within Allomember E of the Dunvegan Formation (see also Figures
45 and 48). Note numerous small deltas associated with the highstand shingles E4 to E3. The youngest shingle, E1, is accompanied
by fluvial entrenchment, incision of an incised valley, and development of a single, massive delta lobe. E1 is thus interpreted as
a lowstand delta. From Bhattacharya (1991).

prodelta shelf plumes (Winn, 1991), to tidally modified shelf stand surfaces of erosion recording the bypass of sediments
ridges (Tillman and Merewether, 1998), to incised estuarine basinward (Bhattacharya and Willis, 2001; Martinsen, 2003). Chert
valley fills (Tillman and Merewether, 1998). A detailed pebble and cobble lags associated with these surfaces are the
allostratigraphic study of the lower portion of the Frontier incor- primary evidence that coarser-grained rivers fed these shorelines
porated lithofacies, ichnofacies, palynofacies, paleocurrent data, (Fig. 55). The fluvial deposits have been completely reworked
bedding relationships, and isolith maps from nearly 100 mea- into a transgressive lag.
sured outcrop sections and about 550 subsurface well logs. Four The low-accommodation setting left little room for sand-
allomembers, interpreted as major delta lobes, were identified stones to stack vertically, and successive episodes of delta
and mapped (Fig. 47). Each allomember has a lobate to elongate progradation were offset along strike (Fig. 47). More tide-influ-
geometry, basinward-dipping internal clinoform bedding (Fig. enced delta deposits of the Frewens Allomember (Willis et al.,
42), radiating paleocurrents, and low to moderate degree of 1999) formed within shoreline embayments defined by the
shallow marine burrowing. Facies successions show variable topography of older wave-influenced delta lobes and subtle
wave- and tide-influenced lithofacies. Delta-plain, paralic, and syndepositional deformation of the basin floor (Bhattacharya
nonmarine facies have been eroded from the top of every deltaic and Willis, 2001).
succession, at least in the Powder River Basin. Erosion surfaces
capping progradational deltaic successions are the only stratal San Miguel Formation
discontinuities that can be mapped regionally (Fig. 55) and were
used to define the four allomembers within the Formation. The In a final example, Weise (1980) used somewhat more tradi-
bounding discontinuities record transgressive ravinement that tional lithostratigraphic and facies techniques to correlate and
was enhanced over areas of structural uplift, rather than low- map sandstones within the Cretaceous San Miguel Formation in
DELTAS 281

FIG. 52.—Block diagram showing paleogeography of delta lobe of


shingle E1, Dunvegan Formation. Allostratigraphic correla-
tions allow channels and associated delta lobes to be sepa-
rated and linked. Compare with Figure 48 and Figure 49
(From Bhattacharya, 1991).

Although the paper was published before the sequence strati-


graphic revolution, Weise suggests that the lack of delta-plain
facies resulted from transgressive reworking and removal of
topsets.
The San Miguel is one of the earliest well-documented ex-
amples of a top-truncated delta system. It illustrates the key point
that the lower the accommodation rate, the greater the degree of
erosion resulting in partial preservation of an ancient deposi-
tional system. These examples show that detailed sedimentol-
ogy, ichnology, facies mapping, and paleocurrent data must all
be integrated to make an unequivocal interpretation, as was
recently elucidated by Martinsen (2003). Although fluvial influ-
ence can be identified in some single vertical successions, many
deltas may contain little evidence of fluvial effects (such as the
San Miguel example), which can then be identified only on the
basis of correlation and mapping.

Controls on Sequence Stratigraphic Organization


The distribution of lithologies and overall facies architecture
within a given depositional system is highly sensitive to allocyclic
controls, such as eustasy, tectonics, and sediment supply (Jervey,
1988). Sequence stratigraphy provides a means of interpreting
these controls on sediment partitioning within and between
different depositional systems, as well as providing a means for
FIG. 51.—Spectrum of sandbody geometries shown by sandstone understanding the origin of key bounding discontinuities that
isolith maps of shingles within various allomembers in the are critical to correlate and map depositional systems and sys-
Upper Cretaceous Dunvegan Formation (Alberta, Canada). tems tracts (Posamentier et al., 1988; Bhattacharya, 1993; Ander-
From Bhattacharya and Walker (1992), based on data pre- son, J.B. et al., 2004). For example, when base level is low or falling
sented by Bhattacharya and Walker (1991b). Mapping of (lowstand and falling-stage (forced regressive) systems tracts),
isoliths within entire allomember does not reveal deltaic river-dominated deltas may form, and the fluvial systems may be
shape as clearly. sandier and may be predominantly erosional or incised
(Bhattacharya and Walker, 1992; Shanley and McCabe, 1994).
During times of lowered sea level in the Cretaceous Western
east Texas, U.S.A., (Fig. 56). Cores showed highly bioturbated Interior, the effects of subtle sea-floor topography may be en-
shelf and shoreface type, upward-coarsening facies successions, hanced, causing complex physiography and numerous
with abundant hummocky cross stratification and a typical wave- embayments of the shoreline (Bhattacharya and Willis, 2001;
dominated Cruziana ichnofacies grading into an Ohiomorpha- Nummedal and Riley, 1999). Tidal effects, in particular, can be
dominated Skolithos ichnofacies. No fluvial or paralic facies were enhanced within these embayments, especially during the initial
noted in the subsurface. Her maps (Fig. 56) showed classic deltaic turnaround of sea level. Units originally interpreted as shelf
morphologies, similar to the models of Coleman and Wright sands, such as the Shannon, Tocito, and Frewens sandstones,
(1975). The sand-body geometries were readily interpreted to be have been reinterpreted as tidally influenced lowstand deltas and
wave-influenced deltas, and the paper has become a classic. shorefaces (Bergman, 1994; Sullivan et al., 1997; Nummedal and
282 JANOK P. BHATTACHARYA

kilometers

FIG. 53.—Allostratigraphy of the Dunvegan Alloformation. Plint (2000) extended correlation of marine fooding surfaces (initiated
by Bhattacharya, 1993) into the nonmarine portion of the clastic wedge. This included mapping of tributive valley systems (see
Figure 38). Nonmarine facies show a distinct thickening, interpreted as driven by increased tectonic subsidence to the
northwest.

Riley, 1999; Willis et al., 1999; Bhattacharya and Willis, 2001; level, and may coincide with development of shoal-water deltas
Willis and Gabel, 2001). in a transgressive or highstand systems tract. The generalizations
In contrast, during rising base level (e.g., transgressive and embodied in the definitions of systems tracts allow the tracts to be
highstand systems tracts) mud tends to be trapped within the used predictively, and in that sense they can be used as basin-
estuary or within the alluvial realm, resulting in more heteroge- scale facies models. One part of a given systems tract may be
neous fluvial reservoirs (see Ferron example, above). Coeval important in predicting the appearance and nature of a related
shorelines may tend to be wave- or tide-dominated (Bhattacharya, part.
1993; Barton, 1997; Plint, 2000). The differences in interpretations and nomenclature in se-
Galloway (1975) suggested that many regressive–transgres- quence stratigraphy results from the distinction between systems
sive clastic wedges in the Texas Gulf Coast show a transition from tracts defined purely on geometric character and physical posi-
elongate river-dominated types in the lower parts of a wedge to tion within the sequence versus definition in the context of time
more wave-modified deltas in the upper parts. Other examples and relative changes in sea level. In most papers on sequence
include the Norias delta system (Duncan, 1983). This idea was stratigraphy, concept-driven genetic models and field observa-
recognized earlier by Barrell (1912), who noted that the Catskill tions are complexly intertwined. Nomenclature problems of the
wedge showed a similar transition. same kind exist in other areas of sedimentology, such as the use
An understanding of the allostratigraphic framework poten- of the term facies to describe rock properties (e.g., cross-bedded
tially allows the geologist to predict relative shale proportions, sandstone facies) versus paleodepositional environments
geometries, and distributions in different depositional environ- (shoreface facies). The way that most scientific systems are named
ments as a function of stratigraphic position and basin setting and analyzed reflects how we think about them, and to some
(e.g., Gardner, 1995). This is of great value in constructing de- degree all observations are model driven. The use of systems
tailed geo-cellular models in reservoir characterization. tracts in the context of relative sea-level change seems to be more
The concept of systems tracts also provides a way of linking widespread than the use favored by Van Wagoner (1995), but in
the different types of depositional systems in various positions in general it is critical to separate observation from interpretations.
the basin. Growth of deep-water submarine fans is commonly
related to the development of shelf-edge deltas in lowstand CONCLUSIONS
systems tracts and the development of hyperpycnal flow at the
river mouth (Por™bski and Steel, 2003; Anderson, 2005; Plink- Deltas are complex three-dimensional progradational depo-
Björklund and Steel, 2005). Development of fans may also corre- sitional systems that form primarily as a consequence of the
late with the incision of alluvial systems in a landward direction. interaction of a river plume with basinal processes, chiefly tides
Submarine-fan growth commonly ends with a relative rise of sea and waves. Longer-term changes in controlling parameters
DELTAS

FIG. 54.—Sequence stratigraphic cross section of the Cretaceous Ferron sandstone. The cross section is based on nearly 100 measured sections and correlations of coals,
bentonites, and other key flooding surfaces in nearly continuous cliff exposures along the Coal Cliffs and Wasatch Plateau in central Utah, U.S.A. (from Garrison and
van den Bergh, 2004).
283
284
JANOK P. BHATTACHARYA

FIG. 55.—Photos of erosional discontinuity above Belle Fourche sandstone bodies of the Frontier Formation, Wyoming, U.S.A.. A) Regional photo shows dipping delta-
front sandstones truncated on top and overlain by shale. B) Close-up of erosional contact between shales and underlying delta-front sandstones. C) Locally, a cross-
bedded pebbly sandstone lag containing sharks’ teeth and oyster fragments marks the transgressive erosion (see Bhattacharya and Willis, 2001, for more details).
DELTAS 285

FIG. 56.—Spectrum of river-dominated (left) to wave-dominated (right) deltas in the Upper Cretaceous San Miguel Formation (Texas,
U.S.A.) after Weise (1980). The isopach maps are contoured in 20 foot (6.1 m) intervals, with maximum thicknesses of about 120–
140 feet (36–43 m). Symmetrical and asymmetrical deltas can be identified. The model for the formation of asymmetrical deltas
suggests that the original interpretations of longshore-drift directions may be reversed from that originally interpreted. Sandier
deposits may occur on the updrift rather than on the downdrift side.

may cause deltas to be transformed into other depositional M.K.G., and Pickering, K.T., eds., Deltas: Sites and Traps for Fossil
systems. A delta may transform into a barrier-island system Fuels: Geological Society of London, Special Publication 41, p. 11–
during a major transgression, for example (Fig. 7). Conversely, 19.
an increase in sedimentation rate, or a channel avulsion, could ALLEN, J.R.L., 1970, Sediments of the modern Niger delta, a summary and
transform a prograding strandplain into a delta. A steady rise of review, in Morgan, J.P., ed., Deltaic Sedimentation; Modern and
relative sea level may cause a river channel or distributary to Ancient: Society of Economic Paleontologists and Mineralogists,
widen into an estuary with tidally influenced sand bars. The Special Publication 15, p. 138–151.
estuary might in turn be drowned, and the system would evolve ALLEN, G.P., LAURIER, D., AND THOUVENENIN, J., 1979, Etude sédimentologique
into a series of shallow-marine, tidal sand ridges (Fig 20; Yang du delta de la Mahakam: Paris, TOTAL, Compagnie Francaise des
and Sun, 1988). Petroles, Notes et Memoires 15, 156 p.
In other settings, waves are important in modifying deltaic ALLISON, M.A., KINEKE, G.C., GORDON, E.S., AND GOÑI, M.A., 2000, Develop-
sediments during transgression. Around the margins of the ment and reworking of a seasonal flood deposit on the inner continen-
Mississippi delta, such modification has produced winnowed tal shelf off the Atchafalaya River: Continental Shelf Research, v. 20,
sandbodies that began as beach-ridge complexes, were detached p. 2267–2294.
as barriers, and have now been drowned to produce shelf shoals ALLISON M.A., KHAN S.R., GOODBRED, S.L., JR., AND KUEHL, S.A., 2003,
(Fig. 7). It is now recognized that waves, rivers, and tides continu- Stratigraphic evolution of the late Holocene Ganges–Brahmaputra
ously interact, resulting in deposits that may show an interfinger- lower delta plain: Sedimentary Geology, v. 155, p. 317–342
ing of facies architectural elements formed by these different ANDERSON, J.B., 2005, Diachronous development of late Quaternary shelf-
processes (Bhattacharya and Giosan, 2003). Depending on the margin deltas in the northwestern Gulf of Mexico: Implications for
larger stratigraphic context, these can be regarded as components sequence stratigraphy and deep-water reservoir occurrence, in Giosan,
of deltaic systems or as distinct depositional systems. L., and Bhattacharya, J.P., eds., River Deltas—Concepts, Models, and
The definition and understanding of a depositional system Examples: SEPM, Special Publication 83, p. 257–276.
depends largely on the scale of observation and objectives of the ANDERSON, J.B., AND FILLON, R.H., eds., 2004, Late Quaternary Stratigraphic
study. The distribution of sediment at the surface in many mod- Evolution of the Northern Gulf of Mexico Margin: SEPM, Special
ern deltas may not characterize the sandbody geometry in the Publication 79, 311 p.
immediately underlying deposits. In many modern deltas, trans- ANDERSON, J.B., RODRIGUEZ, A., ABDULAH, K., FILLON, R.H., BANFIELD, L.A.,
gressive reworking by waves and tides of the uppermost veneer MCKEOWN, H.A., AND WELLNER, J.S., 2004, Late Quaternary strati-
of sediments (i.e., destructional phase) may not reflect the sedi- graphic evolution of the Northern Gulf of Mexico Margin: A synthe-
mentary processes in the major proportion of the underlying sis, in Anderson, J.B., and Fillon, R.H., eds., Late Quaternary Strati-
regressive deltaic package (i.e., constructional phase). graphic Evolution of the Northern Gulf of Mexico Margin: SEPM,
Special Publication 79, p. 1–23.
REFERENCES ANDERSON, P.B., CHIDSEY, T.C., RYER, T.C., ADAMS, R.D., AND MCCLURE, K.
2004, Geologic framework, facies, paleogeography, and reservoir ana-
AINSWORTH, R.B., SANLUNG, M., THEO, S., AND DUIVENVOORDEN, C., 1999, logs of the Ferron sandstone in the Ivie Creek area, East-Central Utah,
Correlation techniques, perforation strategies, and recovery factors: in Chidsey, T.C., Jr., Adams, R.D., and Morris, T.H., eds., The Fluvial–
An integrated 3-D reservoir modeling approach Sirkit Field, Thai- Deltaic Ferron Sandstone: Regional to Wellbore-Scale Outcrop Analog
land: American Association of Petroleum Geologists, Bulletin, v. 83, Studies and Application to Reservoir Modeling: American Association
p. 535–1551. of Petroleum Geologists, Studies in Geology, no. 50, p. 331–356.
AINSWORTH, R.B., AND PATTISON, S.A.J., 1994, Where have all the lowstands ARNOTT, R.W.C., 1992, The role of fluvial processes during deposition of
gone? Evidence for attached lowstand systems tracts in the Western the (Cardium) Carrot Creek/Cyn-Pem conglomerates: Bulletin of
Interior of North America: Geology, v. 22, p. 415–418. Canadian Petroleum Geology, v. 40, p. 356–362.
ALEXANDER, J., 1989, Deltas or coastal plain? With an example of the AUGUSTINIUS, P.G.E.F., 1989, Cheniers and chenier plains: a general intro-
controversy from the Middle Jurassic of Yorkshire, in Whateley, duction: Marine Geology, v. 90, p. 219–230.
286 JANOK P. BHATTACHARYA

BAKER, E.K., HARRIS, P.T., KEENE, J.B., AND SHORT, S.A., 1995, Patterns of The Ferron Sandstone of Utah: American Association of Petroleum
sedimentation in the macrotidal Fly River Delta, Papua, New Guinea, Geologists, Studies in Geology, no. 50, p. 279–304.
in Flemming, B.W., and Bartholomae, A., eds., Tidal Signatures in BHATTACHARYA, J.P., AND DAVIES, R.K., 2001, Growth faults at the prodelta
Modern and Ancient Sediments: International Association of Sedi- to delta front transition, Cretaceous Ferron Sandstone, Utah: Marine
mentologists, Special Publication 24, p. 193–211. and Petroleum Geology, v. 18, p. 525–534.
BARRELL, J., 1912, Criteria for the recognition of ancient delta deposits: BHATTACHARYA, J.P., AND GIOSAN, L., 2003, Wave-influenced deltas: geo-
Geological Society of America, Bulletin, v. 23, p. 377–446. morphological implications for facies reconstruction: Sedimentol-
BART, P.J., AND ANDERSON, J.B., 2004, Late Quaternary stratigraphic evolu- ogy, v. 50, p. 187–210.
tion of the Alabama and west Florida outer continental shelf, in BHATTACHARYA, J.P., AND POSAMENTIER, H.W., 1994, Sequence stratigraphic
Anderson, J.B., and Fillon, R.H., eds., Late Quaternary Stratigraphic and allostratigraphic applications in the Alberta Foreland Basin, in
Evolution of the Northern Gulf of Mexico Margin: SEPM, Special Mossop, G.D., and Shetsen, I., compilers, Geological Atlas of the
Publication 79, p. 43–53. Western Canada Sedimentary Basin: Canadian Society of Petroleum
BARTON, M.D., ANGLE, E.S., AND TYLER, N., 2004, Stratigraphic architec- Geologists and Alberta Research Council, p. 407–412.
ture of fluvial–deltaic sandstones from the Ferron Sandstone out- BHATTACHARYA, J.P., AND TYE, R.S., 2004, Searching for modern Ferron
crop, East-Central Utah, in Chidsey, T.C., Adams, R.D., and Mor- analogs and application to subsurface interpretation, in Chidsey,
ris, T.H., eds., Regional to Wellbore Analog for Fluvial–Deltaic T.C., Jr., Adams, R.D., and Morris, T.H., eds., The Fluvial-deltaic
Reservoir Modeling: The Ferron Sandstone of Utah: American Ferron Sandstone: Regional to Wellbore-Scale Outcrop Analog Stud-
Association of Petroleum Geologists, Studies in Geology, no. 50, p. ies and Application to Reservoir Modeling: American Association of
193–210. Petroleum Geologists, Studies in Geology, no. 50, p. 39–57.
BARTON, M., 1997, Application of Cretaceous Interior Seaway outcrop BHATTACHARYA, J., AND WALKER, R.G., 1991a, Allostratigraphic subdivision
investigations to fluvial–deltaic reservoir characterization: Part 1, of the Upper Cretaceous Dunvegan, Shaftesbury, and Kaskapau
predicting reservoir heterogeneity in delta front sandstones, Ferron Formations in the subsurface of northwestern Alberta: Bulletin of
gas field, Central Utah, in Shanley, K.W., and Perkins, B.E., eds., Canadian Petroleum Geology, v. 39, p. 145–164.
Shallow Marine and Nonmarine Reservoirs, Sequence Stratigraphy, BHATTACHARYA, J., AND WALKER, R.G., 1991b, Facies and facies successions
Reservoir Architecture and Production Characteristics: Gulf Coast in river- and wave-dominated depositional systems of the Upper
Section. SEPM Foundation, Eighteenth Annual Research Confer- Cretaceous Dunvegan Formation, northwestern Alberta: Bulletin of
ence, p. 33–40. Canadian Petroleum Geology, v. 39, p. 165–191.
BATES, C.D., 1953, Rational theory of delta formation: American Associa- BHATTACHARYA, J.P., AND WALKER, R.G., 1992, Deltas, in Walker, R.G., and
tion of Petroleum Geologists, Bulletin, v. 37, p. 2119–2162. James, N.P., eds., Facies Models: Response to Sea Level Change:
BERG, O.R., 1982, Seismic detection and evaluation of delta and turbidite Geological Association of Canada, p. 157–177.
sequences: their application to exploration for the subtle trap: Ameri- BHATTACHARYA, J.P., AND WILLIS, B.J., 2001, Lowstand deltas in the Fron-
can Association of Petroleum Geologists, Bulletin, v. 66, p. 1271–1288. tier Formation, Powder River Basin, Wyoming: Implications for
BERG, R.B., AND AVERY, A.H., 1995, Sealing properties of Tertiary growth sequence stratigraphic models: American Association of Petroleum
faults, Texas Gulf coast: American Association of Petroleum Geolo- Geologists, Bulletin, v. 85, p. 261–294.
gists, Bulletin, v. 79, p. 375–393. BLUM, M.D., AND TÖRNQVIST, T.E., 2000, Fluvial responses to climate and
BERGMAN, K.M., 1994, Shannon sandstone in Hartzog Draw–Heldt Draw sea-level change: a review and look forward: Sedimentology, v. 47, p.
fields reinterpreted as detached lowstand shoreface deposits: Journal 2–48.
of Sedimentary Research v. B64, p. 184–201. BOWEN, D.W., AND WEIMER, P., 2003, Regional sequence stratigraphic
BHATTACHARYA, J., 1989, Estuarine channel fills in the Upper Cretaceous setting and reservoir geology of Morrow incised-valley sand-
Dunvegan Formation: core example, in Reinson, G.E., ed., Modern stones (lower Pennsylvanian), eastern Colorado and western Kan-
and Ancient Examples of Clastic Tidal Deposits—A Core and Peel sas: American Association of Petroleum Geologists, Bulletin, v. 87,
Workshop: Canadian Society of Petroleum Geologists, Calgary, p. 781–815.
Alberta, p. 37–49. BOYD, R., AND PENLAND. S., 1988, A geomorphic model for Mississippi delta
BHATTACHARYA, J.P., 1991, Regional to subregional facies architecture of evolution: Gulf Coast Association of Geological Societies, Transac-
river-dominated deltas in the Alberta subsurface, Upper Cretaceous tions, v. 38, p. 443–452.
Dunvegan Formation, in Miall, A.D., and Tyler, N., eds., The Three- BOYD, R., SUTER, J., AND PENLAND, S., 1989, Sequence stratigraphy of the
Dimensional Facies Architecture of Terrigenous Clastic Sediments, Mississippi delta: Gulf Coast Association of Geological Societies,
and Its Implications for Hydrocarbon Discovery and Recovery: Transactions, v. 39, p. 331–340.
SEPM, Concepts and Models in Sedimentology and Paleontology, BRIDGE, J.S., 2000, The geometry, flow patterns and sedimentary processes
no. 3, p. 189–206. of Devonian rivers and coasts, New York and Pennsylvania, USA, in
BHATTACHARYA, J.P., 1993, The expression and interpretation of marine Friend, P.F., and Williams, B.P.J., eds., New Perspectives on the Old
flooding surfaces and erosional surfaces in core: examples from the Red Sandstone: Geological Society of London, Special Publication
Upper Cretaceous Dunvegan Formation in the Alberta foreland 180, p. 85–108
basin, in Summerhayes, C.P., and Posamentier, H.W., eds., Sequence BRIDGE, J.S., 2003, Rivers and Floodplains: London, Blackwell, 491 p.
Stratigraphy and Facies Associations: International Association of BRISTOW, C.S., AND BEST, J.L., 1993, Braided rivers: perspectives and prob-
Sedimentologists, Special Publication 18, p. 125–160. lems, in, Best, J.L., and Bristow C.S., eds., Braided Rivers: Geological
BHATTACHARYA, J.P., 1994, Cretaceous Dunvegan Formation strata of the Society of London, Special Publication 75, p. 1–11.
Western Canada Sedimentary Basin, in Mossop, G.D., and Shetsen, I., BROWN, A.R., 2005, Interpretation of Three-Dimensional Seismic Data, 6th
compilers, Geological Atlas of the Western Canada Sedimentary Edition: American Association of Petroleum Geologists, Memoir 42,
Basin: Canadian Society of Petroleum Geologists and Alberta Re- 541 p.
search Council, p. 365–373. BROWN, L.F., AND FISHER, W.L., 1977, Seismic-stratigraphic interpretation
BHATTACHARYA, J.P., AND DAVIES, R.K., 2004, Sedimentology and structure of depositional systems: examples from Brazilian rift and pull-apart
of growth faults at the base of the Ferron Member along Muddy basins, in Payton, C.E., ed., Seismic Stratigraphy—Applications to
Creek, Utah, in Chidsey, T.C., Adams, R.D., and Morris, T.H., eds., Hydrocarbon Exploration: American Association of Petroleum Ge-
Regional to Wellbore Analog for Fluvial–Deltaic Reservoir Modeling: ologists, Memoir 26, p. 213–248.
DELTAS 287

BRUUN, P., 1962, Sea level rise as a cause of erosion: American Society of eds., Coarse-Grained Deltas: International Association of Sedimen-
Civil Engineers, Proceedings, Journal of the Waterways and Harbors tologists, Special Publication 10, p. 155–168.
Division, v. 88, no. WW 1, p. 117–130. COWAN, C.A., AND JAMES, N.P., 1992, Diastasis cracks: mechanically gener-
BURK, C.F., JR., 1963, Structure, isopach, and facies maps of Upper Creta- ated syneresis-like cracks in Upper Cambrian shallow water oolite
ceous marine successions, West-Central Alberta and Adjacent British and ribbon carbonates: Sedimentology, v. 39, p. 1101–1118.
Columbia: Geological Survey of Canada, Paper 62-31, 10 p. DALRYMPLE, R.W., 1992, Tidal depositional systems, in Walker, R.G., and
BURNS, B.A., HELLER, P.L., MARZO, M., AND PAOLA, C., 1997, Fluvial response James, N.P., eds., Facies Models: Response to Sea Level Change: St.
in a sequence stratigraphic framework: Example for the Montserrat John’s, Newfoundland, Canada, Geological Association of Canada,
Fan Delta, Spain: Journal of Sedimentary Research, v. 67, p. 311–321. p. 195–218.
BUSCH, D.A., 1971, Genetic units in delta prospecting: American Associa- DALRYMPLE, R.W., 1999, Tide-dominated deltas, do they exist or are they all
tion of Petroleum Geologists, Bulletin, v. 55, p. 1137–1154. estuaries? (abstract): American Association of Petroleum Geologists,
BUSCH, D.A., 1974, Stratigraphic Traps in Sandstones—Exploration Tech- Annual Convention, San Antonio, Texas, Official Program with Ab-
niques: American Association of Petroleum Geologists, Memoir 21, stracts, p. A29.
174 p. DALRYMPLE, R.W., ZAITLIN, B.A., AND BOYD, R., 1992, Estuarine facies mod-
CARBONEL, P., AND MOYES, J., 1987, Late Quaternary paleoenvironments of els: conceptual basis and stratigraphic implications: Journal of Sedi-
the Mahakham delta (Kalimantan, Indonesia): Palaeogeography, mentary Petrology, v. 62, p. 1130–1146.
Palaeoclimatology, Palaeoecology, v. 61, p. 265–284. DALRYMPLE, R.W. BAKER, E.K. HARRIS, P.T., AND HUGHES, M.G., 2003, Sedi-
CATTANEO, A., CORREGGIARI, A., LANGONE, L., AND TRINCARDI, F., 2003, The mentology and stratigraphy of a tide-dominated, foreland-basin
late-Holocene Gargano subaqueous delta, Adriatic shelf: Sediment delta (Fly River, Papua New Guinea), in Sidi, F.H., Nummedal, D.,
pathways and supply fluctuations: Marine Geology, v. 193, p. 61–91. Imbert, P., Darman, H., and Posamentier, H.W., eds., Tropical Deltas
CHIDSEY, T.C., ADAMS, R.D., AND MORRIS T.H., EDS., 2004, Regional to of Southeast Asia—Sedimentology, Stratigraphy, and Petroleum
Wellbore Analog for Fluvial–Deltaic Reservoir Modeling: The Ferron Geology: SEPM, Special Publication 76, p. 147–173
Sandstone of Utah: American Association of Petroleum Geologists, DAVIES, C., BEST, J. , AND COLLIER, R., 2003, Sedimentology of the Bengal
Studies in Geology, no. 50, 568 p. shelf, Bangladesh: comparison of late Miocene sediments, Sitakund
CLEAVES, A.W., AND BROUSSARD, M.C., 1980, Chester and Pottsville deposi- anticline, with the modern, tidally dominated shelf: Sedimentary
tional systems, outcrop and subsurface, in the Black Warrior Basin of Geology, v. 155, p. 271–300.
Mississippi and Alabama: Gulf Coast Association of Geological Soci- DOMINGUEZ, J.M.L., 1996, The São Francisco strandplain: a paradigm for
eties, Transactions, v. 30, p. 49–60. wave-dominated deltas? in de Batist, M., and Jacobs, P., eds., Geol-
COLEMAN, J.M., AND GAGLIANO, S.M., 1964, Cyclic sedimentation in the ogy of Siliciclastic Shelf Seas: Geological Society of London, Special
Mississippi River deltaic plain: Gulf Coast Association of Geological Publication 117, p. 217–231.
Societies, Transactions, v. 14, p. 67–80. DOMINGUEZ, J.M.L., MARTIN, L., AND BITTENCOURT, A.C.S.P., 1987, Sea-level
COLEMAN, J.M., ROBERTS, H.H., MURRAY, S.P., AND SALAMA, M., 1981, Mor- history and Quaternary evolution of river mouth-associated beach-
phology and dynamic stratigraphy of the eastern Nile shelf: Marine ridge plains along the east-southeast Brazilian Coast: a summary, in
Geology, v. 42, p. 301–326. Nummedal, D., Pilkey, O.H., and Howard, J.D., eds., Sea-Level
COLEMAN, J.M., AND PRIOR, D.B., 1982, Deltaic environments, in Scholle, Fluctuation and Coastal Evolution: SEPM, Special Publication 41, p.
P.A., and Spearing, D.R., eds., Sandstone Depositional Environ- 115–127.
ments: American Association of Petroleum Geologists, Memoir 31, p. DUNCAN, E.A., 1983, Delineation of delta types: Norias delta system, Frio
139–178. Formation, South Texas: Gulf Coast Association of Geological Societ-
COLEMAN, J.M., PRIOR, D.B., AND LINDSAY, J.F., 1983, Deltaic influences on ies, Transactions, v. 33, p. 269–273.
shelf edge instability processes, in Stanley, D.J., and Moore, G.T., eds., DRAUT, A.E., KINEKE, G.C., VELASCO, D.W., ALLISON, M.A., AND PRIME, R.J.,
The Shelfbreak: Critical Interface on Continental Margins: SEPM, 2005, Influence of the Atchafalaya River on recent evolution of the
Special Publication 33, p. 121–137. chenier-plain inner continental shelf, northern Gulf of Mexico: Con-
COLEMAN, J.M., AND WRIGHT, L.D., 1975, Modern river deltas: variability of tinental Shelf Research, v. 25, p. 91–112.
processes and sand bodies, in Broussard, M.L., ed., Deltas, Models for EDWARDS, M.B., 1981, Upper Wilcox Rosita delta system of South Texas:
Exploration: Houston, Houston Geological Society, p. 99–149 growth-faulted shelf-edge deltas: American Association of Petro-
COLELLA, A., AND PRIOR, D.B., 1990, Coarse-Grained Deltas: International leum Geologists, Bulletin, v. 65, p. 54–73.
Association of Sedimentologists, Special Publication 10, 357 p. ELLIOTT, T., 1974, Interdistributary bay sequences and their genesis: Sedi-
CORBEANU, R.M., WIZEVICH, M.C., BHATTACHARYA, J.P., ZENG, X., AND mentology, v. 21, p. 611–622.
MCMECHAN, G.A., 2004, Three-dimensional architecture of ancient ELLIOTT, T., 1975, The sedimentary history of a delta lobe from a Yoredale
lower delta-plain point bars using ground-penetrating radar, Cre- (Carboniferous) cyclothem: Yorkshire Geological Society, Proceed-
taceous Ferron Sandstone, Utah, in Chidsey, T.C., Adams, R.D., ings, v. 40, p. 505–536.
and Morris, T.H., eds., Regional to Wellbore Analog for Fluvial– ELLIOTT, T., 1986, Deltas, in Reading, H.G., ed., Sedimentary Environments
Deltaic Reservoir Modeling: The Ferron Sandstone of Utah: Ameri- and Facies: Oxford, U.K., Blackwell Scientific Publications, p. 113–154.
can Association of Petroleum Geologists, Studies in Geology, no. ERIKSSON, K.A., 1979, Marginal marine processes from the Archean Moodies
50, p. 285–309. Group, Barberton Mountain Land, South Africa: evidence and signifi-
CORREGGIARI, A., TRINCARDI, F., LANAGONE, L., AND ROVERI, M., 2001, Styles cance: Precambrian Research, v. 8, p. 153–182.
of failure in heavily sedimented highstand prodelta wedges on the EVAMY, D.D., HAREMBOURE, J., KAMERLING, P., KNAPP, W.A., MOLLOY, F.A.,
Adriatic shelf: Journal of Sedimentary Research, v. 71, p. 218–236. AND ROWLANDS, P.H., 1978, Hydrocarbon habitat of Tertiary Niger
CORREGGIARI, A., CATTANEO, A., AND TRINCARDI, F., 2005, Depositional delta: American Association of Petroleum Geologists, Bulletin, v. 62,
patterns in the Late-Holocene Po Delta system, in Giosan, L., and p. 1–39.
Bhattacharya, J.P., eds., River Deltas—Concepts, Models, and Ex- EYLES, N., AND EYLES, C.H., 1992, Glacial depositional systems, in Walker,
amples: SEPM, Special Publication 83, p. 365–392 R.G., and James, N.P., eds., Facies Models: Response to Sea Level
CORNER, G.D., NORDAHL, E., MUNCH-ELLINGSEN, K., AND ROBERTSON, K.R., Change: Geological Association of Canada, p. 73–100.
1990, Morphology and sedimentology of an emergent fjord-head FIELDING, C.R., TRUEMAN, J., AND ALEXANDER, J., 2005a, Sedimentology of the
Gilbert-type delta: Alta delta, Norway, in Colella, A., and Prior, D.B., modern and Holocene Burdekin River delta of North Queensland,
288 JANOK P. BHATTACHARYA

Australia—Controlled by river output, not by waves and tides, in GILBERT, G.K., 1885, The topographic features of lake shores: U.S. Geologi-
Giosan, L., and Bhattacharya, J.P., eds., River Deltas—Concepts, cal Survey, 5th Annual Report (1883–1884), p. 69–123.
Models, and Examples: SEPM, Special Publication 83, p. 467–496. GINGRAS, M.K., MACEACHERN, J.A., AND PEMBERTON, S.G., 1998, A compara-
FIELDING, C.R., TRUEMAN, J., AND ALEXANDER, J., 2005b, Sharp-based mouth tive analysis of the ichnology of wave- and river-dominated
bar sands from the Burdekin River Delta of northeastern Australia: allomembers of the Upper Cretaceous Dunvegan Formation: Bulletin
extending the spectrum of mouth bar facies, geometry, and stacking of Canadian Petroleum Geology, v. 46, p. 51–73.
patterns: Journal of Sedimentary Research. v. 75, p. 55–66. GIOSAN, L., AND BHATTACHARYA, J.P., 2005, New directions in delta research,
FISHER, W.L., BROWN, L.F., SCOTT, A.J., AND MCGOWEN, J.H., 1969, Delta in Giosan, L., and Bhattacharya, J.P., eds., River Deltas—Concepts,
systems in the exploration for oil and gas, a research colloquium: Models, and Examples: SEPM, Special Publication 83, p. 3–10.
Austin, Texas, Texas Bureau of Economic Geology, 204 p. GOODBRED, S.L., AND KUEHL, S.A., 1999, Holocene and modern sediment
FISK, H.N., 1961, Bar finger sands of the Mississippi delta, in Peterson, J.A., budgets for the Ganges–Brahmaputra River: Evidence for highstand
and Osmond, J.C., eds., Geometry of Sandstone Bodies—A Sympo- dispersal to flood-plain, shelf, and deep-sea depocenters: Geology, v.
sium: Tulsa, American Association of Petroleum Geologists, p. 29–52. 27, p. 559–562.
FRAZIER, D.E., 1974, Depositional episodes: their relationship to the Qua- HALDORSEN, H.H., AND DAMSLETH, E., 1993, Challenges in reservoir charac-
ternary stratigraphic framework in the northwestern portion of the terization: American Association of Petroleum Geologists, Bulletin, v.
Gulf basin: Austin, Texas, Bureau of Economic Geology, Geological 77, p. 541–551.
Circular 74-1, 28 p. HAMPSON, G.J., 2000, Discontinuity surfaces, clinoforms, and facies archi-
GALLOWAY, W.E., 1975, Process framework for describing the morphologic tecture in a wave-dominated, shoreface–shelf parasequence: Journal
and stratigraphic evolution of deltaic depositional systems, in of Sedimentary Research, v. 70, p. 325–340.
Broussard, M.L., ed., Deltas, Models for Exploration: Houston, Texas, HAMPSON, G.J., AND HOWELL, J.A., 2005, Sedimentologic and geomorphic
Houston Geological Society, p. 87–98. characterization of ancient wave-dominated deltaic shorelines: Up-
GALLOWAY, W.E., 1989a, Genetic stratigraphic sequences in basin analysis per Cretaceous Blackhawk Formation, Book Cliffs, Utah, U.S.A., in
I: architecture and genesis of flooding-surface bounded depositional Giosan, L., and Bhattacharya, J.P., eds., River Deltas—Concepts,
units: American Association of Petroleum Geologists, Bulletin, v. 73, Models, and Examples: SEPM, Special Publication 83, p. 133–154.
p. 125–142. HARRIS, P.T., 1988, Large-scale bedforms as indicators of mutually evasive
GALLOWAY, W.E., 1989b, Genetic stratigraphic sequences in basin analysis sand transport and the sequential infilling of wide-mouthed estuar-
II: application to northwest Gulf of Mexico Cenozoic basin: American ies: Sedimentary Geology, v. 57, p. 273–298.
Association of Petroleum Geologists, Bulletin, v. 73, p. 143–154. HART, B.S., AND LONG, B.F., 1996, Forced regressions and lowstand deltas:
GALLOWAY, W.E., AND HOBDAY, D.K., 1996, Terrigenous Clastic Deposi- Holocene Canadian examples: Journal of Sedimentary Research, v.
tional Systems: Heidelberg, Springer-Verlag, 489 p. 66, p. 820–829.
GANI, M.R., AND BHATTACHARYA, J.P., 2005, Bedding correlation vs. facies HAYES, M.O., 1979, Barrier island morphology as a function of tidal and
correlation in deltas: Lessons for Quaternary stratigraphy, in Giosan, wave regime, in Leatherman, S.P., ed., Barrier Islands from the Gulf
L., and Bhattacharya, J.P., eds., River Deltas—Concepts, Models, and of St. Lawrence to the Gulf of Mexico: New York, Academic Press,
Examples: SEPM, Special Publication 83, p. 31–47. p. 1–27.
GARDNER, M.H., 1995, Tectonic and eustatic controls on the stratal archi- HELLAND-HANSEN, W., AND GJELBERG, J.G., 1994, Conceptual basis and
tecture of mid-Cretaceous stratigraphic sequences, central western variability in sequence stratigraphy: a different perspective: Sedi-
interior foreland basin of North America, in Dorobek S.L., and Ross, mentary Geology, v. 92, p. 31–52.
G.M., eds., Stratigraphic Evolution of Foreland Basins: SEPM, Special HISCOTT, R.N., 2003, Latest Quaternary Baram prodelta, northwestern
Publication 52, p. 243–281. Borneo, in Hasan Sidi, F.H., Nummedal, D., Imbert, P., Darman, H.,
GARDNER, M.H., CROSS, T.A., AND LEVORSEN, M., 2004, Stacking patterns, and Posamantier, H.W., eds., Tropical Deltas of Southeast Asia—
sediment volume partitioning and facies differentiation in shallow- Sedimentology, Stratigraphy, and Petroleum Geology: SEPM, Spe-
marine and coastal-plain strata of the Cretaceous Ferron Sandstone, cial Publication 76, p. 89–107.
East-Central Utah, in Chidsey, T.C., Adams, R.D., and Morris, T.H., HOLBROOK, J.M., 1996, Complex fluvial response to low gradients at
eds., Regional to Wellbore Analog for Fluvial–Deltaic Reservoir maximum regression: a genetic link between smooth sequence-bound-
Modeling: The Ferron Sandstone of Utah: American Association of ary morphology and architecture of overlying sheet sandstone: Jour-
Petroleum Geologists, Studies in Geology, no. 50, p. 201–230. nal of Sedimentary Research, v. 66, p. 713–722.
GARRISON, J.R., JR., AND VAN DEN BERGH, T.C.V., 1997, Coal zone and high- HORI, K., SAITO, Y., ZHAO, Q., AND WANG, P., 2002, Architecture and
resolution depositional sequence stratigraphy of the Upper Ferron evolution of the tide-dominated Changjiang (Yangtze) River delta,
Sandstone, in Link, P.K., and Kowallis, B.J., eds., Mesozoic to Recent China: Sedimentary Geology, v. 146, p. 249–264.
Geology of Utah: Brigham Young University, Geology Studies, 42, HORNE, J.C., FERM, J.C., CARUCCIO, F.T., AND BAGANZ, B.P., 1978, Deposi-
Part II, p. 160–178. tional models in coal exploration and mine planning in Appalachian
GARRISON, J.R., JR., AND VAN DEN BERGH, T.C.V., 2004, The high-resolution region: American Association of Petroleum Geologists, Bulletin, v. 62,
depositional sequence stratigraphy of the Upper Ferron Sandstone p. 2379–2411.
Last Chance Delta: an application of coal zone stratigraphy, in Chidsey, JENNETTE, D.C., AND JONES, C.R., 1995, Sequence stratigraphy of the Upper
T.C., Adams, R.D., and Morris, T.H., eds., Regional to Wellbore Cretaceous Tocito Sandstone: a model for tidally influenced incised
Analog for Fluvial–Deltaic Reservoir Modeling: The Ferron Sand- valleys, San Juan basin, Mexico, in Van Wagoner, J.C., and Bertram,
stone of Utah: American Association of Petroleum Geologists, Stud- G.T., eds., Sequence Stratigraphy of Foreland Basin Deposits: Out-
ies in Geology, no. 50, p. 125–192. crop and Subsurface Examples from the Cretaceous of North America:
GASTALDO, R.A., ALLEN, G.P., AND HUC, A.-Y., 1995, The tidal character of American Association of Petroleum Geologists, Memoir 64, p. 311–
fluvial sediments of the modern Mahakam river delta, Kalimantan, 347.
Indonesia, in Flemming, B.W., and Bartholomae, A., eds., Tidal Signa- JERVEY, M.T., 1988, Quantitative geological modeling of siliciclastic rock
tures in Modern and Ancient Sediments: International Association of sequences and their seismic expression, in Wilgus, C.K., Hastings,
Sedimentologists, Special Publication 24, p. 171–181. B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van
GASTESCU, P., 1992, The Danube Delta, Map, Marta Turistica, Compania de Wagoner, J.C., eds., Sea-Level Change: An Integrated Approach:
Turism Pentru Tenuret, Bucharest, Romania. SEPM, Special Publication 42, p. 47–69.
DELTAS 289

JOL, H.M., AND SMITH, D.G., 1991, Ground penetrating radar of northern MACKEY, S.D., AND BRIDGE, J.S., 1995, Three dimensional model of alluvial
lacustrine deltas: Canadian Journal of Earth Sciences, v. 28, p. 1939– stratigraphy: theory and application: Journal of Sedimentary Re-
1947. search, v. B65, p. 7–31.
JOL, H.M., AND SMITH, D.G., 1992, Geometry and structure of deltas in MACNAUGHTON, R.B., DALRYMPLE, R.W., AND NARBONNE, G.M., 1997, Early
large lakes: a ground penetrating radar overview, in Hanninen, P., Cambrian braid-delta deposits, Mackenzie Mountains, north-west-
and Autio, S., eds., Fourth International Conference on Ground ern Canada: Sedimentology, v. 44, p. 587–609.
Penetrating Radar: Geological Survey of Finland, Special Paper 16, MAGUREGUI, J., AND TYLER, N., 1991, Evolution of Middle Eocene tide-
p. 159–168. dominated deltaic sandstones, Lagunillas Field, Maracaibo Basin,
KINEKE, G.C., STERNBERG, R.W., TROWBRIDGE, J.H., AND GEYSER, W.R., 1996, western Venezuela, in Miall, A.D., and Tyler, N., eds., The Three-
Fluid-mud processes on the Amazon continental shelf: Continental Dimensional Facies Architecture of Terrigenous Clastic Sediments,
Shelf Research, v. 16, p. 667–696. and Its Implications for Hydrocarbon Discovery and Recovery: SEPM,
KINEKE, G.C., WOOLFE, K.J., KUEHL, S.A., MILLIMAN, J.D., DELLAPENNA T.M., Concepts and Models in Sedimentology and Paleontology, no. 3, p.
AND PURDON, R.G., 2000, Sediment export from the Sepik River, Papua 233–244.
New Guinea: Evidence for a divergent dispersal system: Continental MARTINSEN, O.J., 1989, Styles of soft-sediment deformation on a Namurian
Shelf Research, v. 20, p. 2239–2266. (Carboniferous) delta slope, Western Irish Namurian basin, Ireland,
KNOX, P.R., AND BARTON, M.D., 1999, Predicting interwell heterogeneity in in Whateley, M.K.G., and Pickering, K.T., eds., Deltas: Sites and Traps
fluvial–deltaic reservoirs: effects of progressive architecture varia- for Fossil Fuels: Geological Society of London, Special Publication 41,
tion through a depositional cycle from outcrop and subsurface obser- p. 167–177.
vations, in Schatzinger, R.A., and Jordan, J.F., eds., Reservoir Charac- MARTINSEN, O.J., 1990, Fluvial, inertia-dominated deltaic deposition in the
terization; Recent Advances: American Association of Petroleum Namurian (Carboniferous) of northern England: Sedimentology, v.
Geologists, Memoir 71, p. 57–72. 37, p. 1099–1113.
KRAFT, J.C., CHRZASTOWSKI, M.J., AND BELKNAP, D.F., 1987, The transgressive MARTINSEN, O.J., 1993, Namurian (late Carboniferous) depositional sys-
barrier–lagoon coast of Delaware: morphostratigraphy, sedimentary tems of the Craven–Askrigg area, northern England: implications for
sequences and responses to relative rise in sea level, in Nummedal, D., sequence stratigraphic models, in Summerhayes, C.P., and Posa-
Pilkey, O.H., and Howard, J.D., eds., Sea-Level Fluctuation and mentier, H.W., eds., Sequence Stratigraphy and Facies Associations:
Coastal Evolution: SEPM, Special Publication 41, p. 129–143. International Association of Sedimentologists, Special Publication 18,
KRAUS, M.J., 2002, Basin-scale changes in floodplain paleosols: implica- p. 247–281.
tions for interpreting alluvial architecture: Journal of Sedimentary MARTINSEN, R.S., 2003, Depositional remnants, part 1: Common compo-
Research, v. 72, p. 500–509 nents of the stratigraphic record with important implications for
KROONENBERG, S., BRUSAKOR, G.V., AND SVITOCH, A.A., 1997, The wandering hydrocarbon exploration and production: American Association of
of the Volga delta: a response to rapid Caspian sea-level changes: Petroleum Geologists, Bulletin, v. 87, p. 1869–1882.
Sedimentary Geology, v. 107, p. 189–209 MATTSON, A., 1997, Characterization, facies relationships, and architec-
KUEHL, S.A., DEMASTER, D.J., AND NITTROUER, C.A., 1986, Distribution of tural framework in a fluvial-deltaic sandstone—Cretaceous Ferron
sedimentary structures on the Amazon subaqueous delta: Continen- Sandstone, central Utah: University of Utah, M.S. thesis, 174 p.
tal Shelf Research, v. 6, p. 311–336. MCCARTHY, P.J., 1999, Evolution of an ancient coastal plain: Palaeosols,
KUEHL, S.A., LEVY, B.M., MOORE, W.S., AND ALLISON, M.A., 1997, Subaque- interfluves and alluvial architecture in a sequence stratigraphic frame-
ous delta of the Ganges–Brahmaputra river system: Marine Geology, work, Cenomanian Dunvegan Formation, NE British Columbia,
v. 144, p. 81–96. Canada: Sedimentology, v. 46, p. 861–891
KUEHL, S.A., ALLISON, M.A., GOODBRED, S.L., AND KUDRASS, H., 2005, The MCPHERSON, J.G., SHANMUGAM, G., AND MOIOLA, R.J., 1987, Fan-deltas and
Ganges–Brahmaputra Delta, in Giosan, L., and Bhattacharya, J.P., braid deltas: varieties of coarse-grained deltas: Geological Society of
eds., River Deltas—Concepts, Models, and Examples: SEPM, Special America, Bulletin, v. 99, p. 331–340.
Publication 83, p. 413–434. MELLERE, D., AND STEEL, R.J., 1995, Facies architecture and sequentiality of
LAMBIASE, J.J., DAMIT, A.R., SIMMONS, M.D., ABDOERRIAS, R., AND HUSSIN, A., nearshore and ‘shelf’ sandbodies: Haystack Mountains Formation,
2003, A depositional model and the stratigraphic development of Wyoming, USA: Sedimentology, v. 42, p. 551–574
modern and ancient tide-dominated deltas in NW Borneo, in Sidi, MELLERE, D., AND STEEL, R.J., 1996, Tidal sedimentation in Inner Hebrides
F.H., Nummedal, D., Imbert, P., Darman, H., and Posamantier, half grabens, Scotland: the Mid-Jurassic Bearreraig Sandstone For-
H.W., eds., Tropical Deltas of Southeast Asia—Sedimentology, mation, in de Batist, M., and Jacobs, P., eds., Geology of Siliciclastic
Stratigraphy, and Petroleum Geology: SEPM, Special Publication Seas: Geological Society of London, Special Publication 117, p. 49–
76, p. 109–123. 79.
LEE, K., ZENG, X., MCMECHAN, G.A., HOWELL, C.D., JR., BHATTACHARYA, MIALL, A.E., 1985, Architectural-element analysis: A new method of facies
J.P., MARCY, F., AND OLARIU, C., 2005, A GPR survey of a delta-front analysis applied to fluvial deposits: Earth-Science Reviews, v. 22, p.
reservoir analog in the Wall Creek Member, Frontier Formation, 308, 1985.
Wyoming: American Association of Petroleum Geologists, Bulle- MIALL, A.D., 1996, The Geology of Fluvial Deposits; Sedimentary Facies,
tin, v. 89, p. 1139–1155. Basin Analysis and Petroleum Geology: Berlin, Springer, 582 p.
LIU, J.P., MILLIMAN, J.D., AND GAO, S., 2002, The Shandong mud wedge and MIALL, A.D., 1997, The Geology of Stratigraphic Sequences: Berlin, Springer,
post-glacial sediment accumulation in the Yellow Sea: Geo-Marine 433 p.
Letters, v. 21, p. 212–218. MICHELS, K.H., KUDRASS, H.R., HUBSCHER, C., SUCKOW, A., AND WIEDICKE, M.,
MACEACHERN, J.A., ZAITLIN, B.A., AND PEMBERTON, S.G., 1998, High-resolu- 1998, The submarine delta of the Ganges–Brahmaputra: cyclone-
tion sequence stratigraphy of early transgressive deposits, Viking dominated sedimentation patterns: Marine Geology, v. 149, p. 133–
Formation, Joffre Field, Alberta, Canada: American Association of 154.
Petroleum Geologists, Bulletin, v. 82, p. 729-755. MONAHAN, P.A., LUTERNAUER, J.L., AND BARRIE, J.V., 1997, The topset and
MACEACHERN, J., BHATTACHARYA, J.P., HOWELL, C.D., AND BANN, K., 2005, foreset of the modern Fraser River Delta, British Columbia, Canada,
Ichnology of deltas, in Giosan, L., and Bhattacharya, J.P., eds., River in Wood, J.M., and Martindale, B., compilers, Sedimentary Events—
Deltas—Concepts, Models, and Examples: SEPM, Special Publication Hydrocarbon Systems, Core Conference, CSPG–SEPM Joint Conven-
83, p. 49–85. tion, p. 491–517.
290 JANOK P. BHATTACHARYA

MOSLOW, T.F., AND PEMBERTON, S.G., 1988, An integrated approach to the mentation, Modern and Ancient: Society of Economic Paleontolo-
sedimentological analysis of some lower Cretaceous shoreface and gists and Mineralogists, Special Publication 15, p. 198–212.
delta front sandstone sequences, in James, D.P., and Leckie, D.A., ORTON, G., AND READING, H.G., 1993, Variability of deltaic processes in
eds., Sequences, Stratigraphy, Sedimentology; Surface and Subsur- terms of sediment supply, with particular emphasis on grain size:
face: Canadian Society of Petroleum Geologists, Memoir 15, p. 373– Sedimentology, v. 40, p. 475–512.
386. OVEREEM, I., KROONENBERG, S.B., VELDKAMP, A., GROENESTEIJN, K., RUSAKOV,
MULDER, T., AND SYVITSKI, J.P.M., 1995, Turbidity currents generated at G.V., AND SVITOCH, A.A., 2003, Small-scale stratigraphy in a large ramp
river mouths during exceptional discharge to the world’s oceans: delta: recent and Holocene sedimentation in the Volga delta, Caspian
Journal of Geology, v. 103, p. 285–298. Sea: Sedimentary Geology, v. 159, p. 133–157.
MULDER, T., SAVOYE, B., SYVITSKI, J.P.M., AND COCHONAT, P., 1996, Origine OVEREEM, I., SYVITSKI, J.P.M., AND HUTTON, E.W.H., 2005, Three-dimen-
des courants de turbidité enregistrés à embouchure du Var en 1971: sional numerical modeling of deltas, in Giosan, L., and Bhattacharya,
Académie des Sciences, Paris, Comptes Rendus, v. 322, Série Iia, p. J.P., eds., River Deltas—Concepts, Models, and Examples: SEPM,
301–307. Special Publication 83, p. 13–30.
MULDER, T., AND ALEXANDER, J., 2001, The physical characteristics of sub- OVEREEM, I., AND WELTJE, G.J., 2001, Conditioning channel switching for a
aqueous sedimentary density flows and their deposits: Sedimentol- 3-D fluvio-deltaic process model: International Association of Math-
ogy, v. 48, p. 269–299. ematical Geologists, 2001, Cancun, Mexico, 6–12 September 2001,
MUTTI, E., ROSELL, J., ALLEN, G.P., FONNESU, F., AND SGAVETTI, M., 1985, The Expanded abstract.
Eocene Baronia tide dominated delta-shelf system in the Ager Basin, PAOLA, C., 2000, Quantitative models of sedimentary basin filling: Sedi-
in Mila, M.D., and Rosell, J., eds., Excursion Guide-book: 6th Euro- mentology, v. 7, p. 121–178.
pean Regional Meeting, International Association of Sedimentolo- PANIN, N., PANIN, S., HERZ, N., AND NOAKES, J.E., 1983. Radiocarbon dating
gists, Lleida, Spain, p. 579–600. of Danube Delta deposits: Quaternary Research, v. 19, p. 249–255.
NEILL, C.F., AND ALLISON, M.A., 2005, Subaqueous deltaic formation on the PARSONS, J.D., BUSH, J.W.M., AND SYVITSKI, J.P.M., 2001. Hyperpycnal plume
Atchafalaya Shelf, Louisiana: Marine Geology, v. 214, p. 411–430. formation from riverine outflows with small sediment concentra-
NELSON, B.W., 1970, Hydrography, sediment dispersal, and recent histori- tions: Sedimentology, v. 48, p. 465–478.
cal development of the Po River delta, Italy, in Morgan, J.P., ed., PAYNE, M.M., GROSVENOR, M.B., GROSVENOR, G.M., PEELE, W.T., COOK, D.W.,
Deltaic Sedimentation; Modern and Ancient: Society of Economic AND BILLARD, J.B., EDS., 1975, National Geographic Atlas of the World,
Paleontologists and Mineralogists, Special Publication 15, p. 152–184. Fourth Edition: Washington, D.C., National Geographic Society, 330 p.
NEMEC, W., 1995, The dynamics of deltaic suspension plumes, in Oti, PENLAND, S., AND SUTER, J., 1989, The geomorphology of the Mississippi
M.N., and Postma, G., eds., Geology of Deltas: Rotterdam, Balkema, River chenier plan: Marine Geology, v. 90, p. 231–258.
p. 31–93. PLINK-BJÖRKLUND, P., AND STEEL, R., 2005, Deltas on falling-stage and
NEMEC, W., STEEL, R.J., GJELBERG, J., COLLINSON, J.D., PRESTHOLM, E., AND lowstand shelf margins, the Eocene Central Basin of Spitsbergen:
OXNEVAD, I.E., 1988, Anatomy of a collapsed and re-established delta Importance of sediment supply, in Giosan, L., and Bhattacharya, J.P.,
front in Lower Cretaceous of Eastern Spitsbergen: gravitational slid- eds., River Deltas—Concepts, Models, and Examples: SEPM, Special
ing and sedimentation processes: American Association of Petroleum Publication 83, p. 179–206.
Geologists, Bulletin, v. 72, p. 454–476. PLINK-BJÖRKLUND, P., AND STEEL, R.J., 2004, Initiation of turbidity currents:
NITTROUER C.A., KUEHL, S.A., DEMASTER, D.J., AND KOWSMANN, R.O., 1986, outcrop evidence for Eocene hyperpycnal flow turbidites: Sedimen-
The deltaic nature of Amazon shelf sedimentation: Geological Society tary Geology, v. 165, p. 29–52.
of America, Bulletin, v. 97, 444–458. PLINT, A.G., 1988, Sharp-based shoreface sequences and “offshore bars” in
NUMMEDAL, D., AND RILEY, G.W., 1999, The origin of the Tocito Sandstone the Cardium Formation of Alberta: Their relationship to relative
and its sequence stratigraphic lessons, in Bergman, K.W., and changes in sea level, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C.,
Snedden, J.W., eds., Isolated Shallow Marine Sand Bodies: Sequence Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea-
Stratigraphic Analysis and Sedimentologic Interpretation: SEPM, Level Changes: An Integrated Approach: SEPM, Special Publication
Special Publication 64, p. 227–254. 42, p. 357–370.
NUMMEDAL, D., AND SWIFT, D.J.P., 1987, Transgressive stratigraphy at PLINT, A.G., 2000, Sequence stratigraphy and paleogeography of a
sequence-bounding unconformities: some principles derived from Cenomanian deltaic complex: the Dunvegan and lower Kaskapau
Holocene and Cretaceous examples, in Nummedal, D., Pilkey, O.H., formations in subsurface and outcrop, Alberta and British Columbia,
and Howard, J.D., eds., Sea-Level Fluctuation and Coastal Evolution: Canada: Bulletin of Canadian Petroleum Geology, v. 47, p. 43–79.
SEPM, Special Publication 41, p. 241–260. PLINT, A.G., AND WADSWORTH, J.A., 2003, Sedimentology and palaeogeo-
NUMMEDAL, D., SIDI, F.H., AND POSAMENTIER, H.W., 2003, A framework for morphology of four large valley systems incising delta plains, West-
deltas in southeast Asia, in Sidi, F.H., Nummedal, D., Imbert, P., ern Canada foreland basin: implications for Mid-Cretaceous sea-level
Darman, H., and Posamantier, H.W., eds., Tropical Deltas of South- changes: Sedimentology, v. 50, p. 1147–1186.
east Asia—Sedimentology, Stratigraphy, and Petroleum Geology: PLUMMER, P.S. AND GOSTIN, V.A., 1981, Shrinkage cracks: desiccation or
SEPM, Special Publication 76, p. 5–17. synaeresis?: Journal of Sedimentary Petrology, v. 51, p. 1147–1156.
OLARIU, C., BHATTACHARYA, J.P., XU, X., AIKEN, C.L.V., ZENG, X., AND POR§BSKI, S.J., AND STEEL, R.J., 2003, Shelf-margin deltas: their stratigraphic
MCMECHAN, G.A., 2005, Integrated study of ancient delta-front significance and relationship to deepwater sands: Earth-Science Re-
deposits, using outcrop, ground-penetrating radar and three-di- views, v. 62, p. 283–326.
mensional photorealistic data: Cretaceous Panther Tongue Sand- POSAMENTIER, H.W., JERVEY, M.T., AND VAIL, P.R., 1988, Eustatic controls on
stone, Utah,, in Giosan, L., and Bhattacharya, J.P., eds., River Del- clastic deposition I—conceptual framework, in Wilgus, C.K., Hastings,
tas—Concepts, Models, and Examples: SEPM, Special Publication B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van
83, p. 155–177. Wagoner, J.C., eds., Sea-Level Changes: An Integrated Approach:
OLARIU, C., AND BHATTACHARYA, J.P., 2006, Terminal distributary channels SEPM, Special Publication 42, p. 109–124.
and delta front architecture of river-dominated delta systems: Jour- POSAMENTIER, H.W., ALLEN, G.P., JAMES, D.P., AND TESSON, M., 1992, Forced
nal of Sedimentary Research, v. 76, p. 212–233. regressions in a sequence stratigraphic framework: concepts, ex-
OOMKENS, E., 1970, Depositional sequences and sand distribution in the amples, and exploration significance: American Association of Petro-
postglacial Rhone delta complex, in Morgan, J.P., ed., Deltaic Sedi- leum Geologists, Bulletin, v. 76, p. 1687–1709.
DELTAS 291

POSAMENTIER, H.W., AND ALLEN, G.P., 1999, Siliciclastic Sequence Stratigra- SCRUTON, P.C., 1960, Delta building and the deltaic sequence, in Shepard,
phy—Concepts and Applications: SEPM, Concepts in Sedimentology F.P., Phleger, F.B., and Van Andel, T.H., eds., Recent Sediments,
and Paleontology, no. 7, 216 p. Northwest Gulf of Mexico: Tulsa, Oklahoma, American Association
POSTMA, G., 1990, Depositional architecture and facies of river and fan of Petroleum Geologists, p. 82–102.
deltas: a synthesis, in Colella. A., and Prior, D.B., eds., Coarse Grained SESTINI, G., 1989, Nile delta: a review of depositional environments and
Deltas: International Association of Sedimentologists, Special Publi- geological history, in Whateley, M.K.G., and Pickering, K.T., eds.,
cation 10, p. 13–27. Deltas: Sites and Traps for Fossil Fuels: Geological Society of London,
PULHAM, A.J., 1989, Controls on internal structure and architecture of Special Publication 41, p. 99–127.
sandstone bodies within Upper Carboniferous fluvial-dominated SHANLEY, K.W., AND MCCABE, P.J., 1994, Perspectives on the sequence
deltas, County Clare, western Ireland, in Whateley, M.K.G., and stratigraphy of continental strata: American Association of Petro-
Pickering, K.T., eds., Deltas: Sites and Traps for Fossil Fuels: Geologi- leum Geologists, Bulletin, v. 78, p. 544–568.
cal Society of London, Special Publication 41, p. 179–203. SHEPARD, F.P., PHLEGER, F.B., AND VAN ANDEL, T.H., EDS., 1960, Recent
RAHMANI, R.A., 1988, Estuarine tidal channel and nearshore sedimenta- Sediments, Northwest Gulf of Mexico: Tulsa, Oklahoma, American
tion of a Late Cretaceous epicontinental sea, Drumheller, Alberta, Association of Petroleum Geologists, 394 p.
Canada, in de Boer, P.L., van Gelder, A., and Nio, S.D., eds., Tide- SIDI, F.H., NUMMEDAL, D., IMBERT, P., DARMAN, H., AND POSAMANTIER, H.W.,
Influenced Sedimentary Environments and Facies: Dordrecht, The EDS., 2003, Tropical Deltas of Southeast Asia—Sedimentology, Stratig-
Netherlands, D. Riedel Publishing Company, p. 433–471. raphy, and Petroleum Geology, SEPM, Special Publication 76, 269 p.
RAMOS, A., AND GALLOWAY, W.E., 1990, Facies and sand-body geometry of SMITH, N.D., PHILLIPS, A.C., AND POWELL, R.D., 1990, Tidal drawdown: a
the Queen City (Eocene) tide-dominated delta-margin embayment, mechanism for producing cyclic laminations in glaciomarine deltas:
NW Gulf of Mexico basin: Sedimentology, v. 37, p. 1079–1098. Geology, v. 18, p. 10–13.
RASMUSSEN, D.L., JUMP, C.J., AND WALLACE, K.A., 1985, Deltaic systems in the SMITH, D.G., JOL, H.M., SMITH, N.D., KOSTASCHUK, R.A., AND PEARCE, C.M.,
Early Cretaceous Fall River Formation, southern Powder River Basin, 2005, Wave-dominated William river delta, its morphology, radar
Wyoming: Wyoming Geological Association, 36th Annual Field Con- stratigraphy, and history, Lake Athabasca, Canada, in Giosan, L., and
ference, Guidebook, p. 91–111. Bhattacharya, J.P., eds., River Deltas—Concepts, Models, and Ex-
READING, H.G., AND COLLINSON, J.D., 1996, Clastic Coasts, in Reading, H.G., amples: SEPM, Special Publication 83, p. 295–318.
ed., Sedimentary Environments; Processes, Facies and Stratigraphy, SNEDDEN, J.W., AND BERGMAN, K.M., 1999, Isolated shallow marine sand
Third Edition: Oxford, U.K., Blackwell Science, p. 154–231. bodies: Deposits for all interpretations, in Bergman, K.M., and Snedden,
REINSON, G.E., 1992, Barriers and estuaries, in Walker, R.G., and James, J.W., eds., Isolated Shallow Marine Sand Bodies: Sequence Strati-
N.P., eds., Facies Models: Response to Sea Level Change: Geological graphic Analysis and Sedimentologic Interpretation: SEPM, Special
Association of Canada, p. 179–194. Publication 64, p. 1–12.
REYNOLDS, A.D., 1999, Dimensions of paralic sandstone bodies: American SORIA, J.M., FERNÁNDEZ, J., GARCÍA, F., AND VISERAS, C., 2003, Correlative
Association of Petroleum Geologists, Bulletin, v. 83, p. 211–229. lowstand deltaic and shelf systems in the Gaudix basin (late Mi-
RICH, J.L., 1951, Three critical environments of deposition and criteria for ocene, Betic Cordillera, Spain): the stratigraphic record of forced
recognition of rocks deposited in each of them: Geological Society of and normal regressions: Journal of Sedimentary Research, v. 73, p.
America, Bulletin, v. 62, p. 1–20. 912–925.
RINE, J.M., AND GINSBURG, R.N., 1985, Depositional facies of a mud STOW, D.A.V., 1986, Deep clastic seas, in Reading, H.G., ed., Sedimentary
shoreface in Suriname, South America: a mud analogue to sandy, Environments and Facies, Second Edition: Oxford, U.K., Blackwell
shallow-marine deposits: Journal of Sedimentary Petrology, v. 55, p. Scientific Publications, p. 399–444.
633–652. SULLIVAN, M.D., VAN WAGONER, J.C., JENNETTE, D.C., FOSTER, M.E., STUART,
ROBERTS, H.H., AND SYDOW, J., 2003, Late Quaternary stratigraphy and R.M., LOVELL, R.W., AND PEMBERTON, S.G., 1997, High resolution
sedimentology of the offshore Mahakham Delta, East Kalimantan sequence stratigraphy and architecture of the Shannon Sandstone,
(Indonesia), in Sidi, F.H., Nummedal, D., Imbert, P., Darman, H., and Hartzog Draw Field, Wyoming: implications for reservoir man-
Posamantier, H.W., eds., Tropical Deltas of Southeast Asia—Sedi- agement, in Shanley, K.W., and Perkins, B.F., eds., Shallow Marine
mentology, Stratigraphy, and Petroleum Geology, SEPM, Special and Nonmarine Reservoirs, Sequence Stratigraphy, Reservoir Ar-
Publication 76, p. 125–145. chitecture and Production Characteristics: Gulf Coast Section,
ROBERTS, H.H, FILLON, R.H., KOHL, B., ROBALIN, J.M., AND SYDOW, J.C., 2004, SEPM Foundation, Eighteenth Annual Research Conference. p.
Depositional architecture of the Lagniappe Delta: sediment charac- 331–344.
teristics, timing of depositional events, and temporal relationship SUTER, J.H., AND BERRYHILL, H.L., JR., 1985, Late Quaternary shelf-margin
with adjacent shelf-edge deltas, in Anderson, J.B., and Fillon, R.H., deltas, Northwest Gulf of Mexico: American Association of Petro-
eds., Late Quaternary Stratigraphic Evolution of the Northern Gulf of leum Geologists, Bulletin, v. 69, p. 77–91.
Mexico Margin: SEPM, Special Publication 79, p. 143–188 SYDOW, J., AND ROBERTS, H.H., 1994, Stratigraphic framework of a late
RODRIGUEZ, A.B., HAMILTON, M.D., AND ANDERSON, J.B., 2000, Facies and Pleistocene shelf-edge delta, northeast Gulf of Mexico: American
evolution of the modern Brazos Delta, Texas: wave versus flood Association Petroleum Geologists, Bulletin, v. 78, p. 1276–1312.
influence: Journal of Sedimentary Research, v. 70, p. 283–295. SYVITSKI, J.P.M., AND MOREHEAD, M.D., 1999, Estimating river-sediment
RYER, T.A., 1984, Transgressive–regressive cycles and the occurrence of discharge to the ocean: application to the Eel margin, Northern
coal in some Upper Cretaceous strata of Utah, U.S.A., in Rahmani, California: Marine Geology, v. 154, p. 13–28.
R.A., and Flores, R.M., eds., Sedimentology of coal and coal-bearing SYVITSKI, J.P.M., HARVEY, N., WOLLANSKI, E., BURNETT, W.C., PERILLO, G.M.E.,
sequences: International Association of Sedimentologists, Special AND GORNITZ, V., 2005, Dynamics of the Coastal Zone, in Crossland,
Publication 7, p. 217–227. C.J., Kremer, H.H., Lindebloom, H.J., Marshall-Crossland, J.I., and Le
SCHEIHING, M.H., AND GAYNOR, G.C., 1991, The shelf sand-plume model: a Tissier, M.D.A., eds., Global Fluxes in the Anthropocene: Berlin,
critique: Sedimentology, v. 38, p. 433–444. Springer, p. 39–94.
SCHUMM, S.A., 1972, Fluvial paleochannels, in Rigby, J.K., and Hamblin, SWIFT. D.J.P., 1968, Coastal erosion and transgressive stratigraphy: Jour-
W.K., eds., Recognition of Ancient Sedimentary Environments: Soci- nal of Geology, v. 76, p. 444–456.
ety of Economic Paleontologists and Mineralogists, Special Publica- TA, T.K.O., NGUYEN, V.L., TATEISHI, M., KOBAYASHI, I., SAITO, Y., AND
tion 16, p. 98–107. NAKAMURA, T., 2002, Sediment facies and Late Holocene progradation
292 JANOK P. BHATTACHARYA

of the Mekong River Delta in Bentre Province, southern Vietnam: an p. 381–399.


example of evolution from a tide-dominated to a tide- and wave- WALKER, R.G., AND PLINT, A.G., 1992, Wave- and storm-dominated shal-
dominated delta: Sedimentary Geology, v. 152, p. 313–325. low marine systems, in Walker, R.G., and James, N.P., eds., Facies
TA, T.K.O., NGUYEN, V.L., TATEISHI, M., KOBAYASHI, I., AND SAITO, Y., 2005, Models: Response to Sea Level Change: Geological Association of
Holocene Delta Evolution and Depositional Models of the Mekong Canada, p. 219–238.
Delta, Southern Vietnam, in Giosan, L., and Bhattacharya, J.P., eds., WARNE, A.G., MEADE, R.H., WHITE, W.A., GUEVARA, E.H., GIBEAUT, J., SMYTH,
River Deltas—Concepts, Models, and Examples: SEPM, Special R.C., ASLAN, A., AND TREMBLAY, T., 2002, Regional controls on geomor-
Publication 83, p. 453–466. phology, hydrology, and ecosystem integrity in the Orinoco Delta,
TESSON, M., ALLEN, G.P., AND RAVÈNNE, C., 1993, Late Pleistocene shelf- Venezuela: Geomorphology, v. 44, p. 273–307.
perched lowstand wedges on the Rhone continental shelf, in WEISE, B.R., 1980, Wave-dominated deltaic systems of the Upper Creta-
Summerhayes, C.P., and Posamentier, H.W., eds., Sequence Stratig- ceous San Miguel Formation, Maverick Basin, South Texas: Austin,
raphy and Facies Associations: International Association of Sedimen- Texas, Texas Bureau of Economic Geology, Report of Investigations
tologists, Special Publication 18, p. 183–196. 107, 39 p.
TILLMAN, R.W., AND MEREWETHER, E.A., 1998, Bayhead-delta and estuary WELDER, F.A., 1959, Processes of Deltaic Sedimentation in the Lower
mouth bars of early Cenomanian age in the mid-Cretaceous Frontier Mississippi River: Alexandria ,Virginia, Defense Technical Informa-
Formation, Central Wyoming (abstract): American Association of tion Center, 90 p.
Petroleum Geologists, Annual Meeting, Extended Abstracts Volume WHITE, C.D., WILLIS, B.J., DUTTON, S.P., BHATTACHARYA, J.P., AND NARAYANAN,
2, A675, p. 1–4. K., 2004, Sedimentology, statistics, and flow behavior for a tide-
TYE, R.S., 2004, Geomorphology: An approach to determining subsurface influenced deltaic sandstone, Frontier Formation, Wyoming, United
reservoir dimensions: American Association of Petroleum Geolo- States, in Grammer, G.M., Harris, P.M., and Eberli, G.P., eds., Integra-
gists, Bulletin, v. 88, p. 1123–1147. tion of Outcrop and Modern Analogs in Reservoir Modeling: Ameri-
TYE, R.S., BHATTACHARYA, J.P., LORSONG, J.A., SINDELAR, S.T., KNOCK, D.G., can Association of Petroleum Geologists, Memoir 80, p. 129–152.
PULS, D.D., AND LEVINSON, R.A., 1999, Geology and stratigraphy of WIGNALL, P.B., AND BEST, J.L., 2004, Sedimentology and kinematics of a
fluvio-deltaic deposits in the Ivishak Formation: Applications for large, retrogressive growth-fault system in Upper Carboniferous
development of Prudhoe Bay Field, Alaska: American Association of deltaic sediments, western Ireland: Sedimentology, v. 52, p. 1343–
Petroleum Geologists, Bulletin, v. 83, p. 1588–1623. 1358.
TYE, R.S., AND COLEMAN, J.M., 1989, Depositional processes and stratigra- WILLIAMS, H.F.L., AND ROBERTS, M.C., 1989, Holocene sea-level change and
phy of fluvially dominated lacustrine deltas: Mississippi Delta Plain: delta growth: Fraser River delta, British Columbia: Canadian Journal
Journal of Sedimentary Petrology, v. 59, p. 973–996. of Earth Sciences, v. 26, p. 1657–1666.
TYLER, N., AND FINLEY, R.J., 1991, Architectural controls on the recovery of WILLIS, B.J., 1997, Architecture of fluvial-dominated valley-fill deposits in
hydrocarbons from sandstone reservoirs, in Miall, A.D., and Tyler, the Cretaceous Fall River Formation: Sedimentology, v. 44, p. 735–
N., eds., The Three-dimensional Facies Architecture of Terrigenous 757.
Clastic Sediments, and Its Implications for Hydrocarbon Discovery WILLIS, B.J., 2005, Tide-influenced river delta deposits, in Giosan, L., and
and Recovery: SEPM, Concepts and Models in Sedimentology and Bhattacharya, J.P., eds., River Deltas—Concepts, Models, and Ex-
Paleontology, no. 3, p. 1–5. amples: SEPM, Special Publication 83, p. 87–129.
ULICNY, D., 2001, Depositional systems and sequence stratigraphy of WILLIS, B.J., BHATTACHARYA, J.B., GABEL., S.L., AND WHITE, C.D, 1999, Archi-
coarse-grained deltas in a shallow-marine, strike-slip setting: the tecture of a tide-influenced delta in the Frontier Formation of Central
Bohemian Cretaceous Basin, Czech Republic: Sedimentology, v. 48, p. Wyoming, USA: Sedimentology, v. 46, p. 667–688.
599–628. WILLIS, B.J., AND GABEL, S., 2001, Sharp-based, tide-dominated deltas of
VAN HEERDEN, I.L., AND ROBERTS, H.H., 1988, Facies development of the Sego Sandstone, Book Cliffs, Utah, USA: Sedimentology, v. 48,
Atchafalaya delta, Louisiana: a modern bayhead delta: American p. 479–506.
Association of Petroleum Geologists, Bulletin, v. 72, p. 439–453. WINKER, C.D., AND EDWARDS, M.B., 1983, Unstable progradational clastic
VAN WAGONER, J.C., MITCHUM, R.M., CAMPION, K.M., AND RAHMANIAN, V.D., shelf margins, in Stanley, D.J., and Moore, G.T., eds., The Shelfbreak:
1990, Siliciclastic Sequence Stratigraphy in Well Logs, Cores, and Critical Interface on Continental Margins: SEPM, Special Publication
Outcrops: American Association of Petroleum Geologists, Methods 33, p. 139–157.
in Exploration Series, no. 7, 55 p. WINN, R.D., 1991, Storm deposition in marine sand sheets: Wall Creek
VONGVISESSOMJAI, S., 1990, Coastal erosion in Thailand: Geographical Member, Frontier Formation, Powder River Basin, Wyoming: Journal
Journal, v. 15, p. 321–337. of Sedimentary Petrology, v. 61, p. 86–101.
VÖRÖSMARTY, C.J., SHARMA, K.P., BALÁZS, M.F., COPELAND, A.H., HOLDEN, J., WOODROW, D.L., AND SEVON, W.D., EDS., 1985, The Catskill Delta: Geological
MARBLE, J., AND LOUGH, J.A., 1997, The storage and aging of continental Society of America, Special Paper 201, 246 p.
runoff in large reservoir systems of the world: Ambio, v. 26, p. 210– WRIGHT, L.D., 1977, Sediment transport and deposition at river mouths: a
219. synthesis: Geological Society of America, Bulletin, v. 88, p. 857–868.
WALKER, R.G., 1984. Facies Models, Second Edition: Geological Associa- WRIGHT, L.D., WISEMAN, W.J., BORNHOLD, B.D., PRIOR, D.B., SUHAYDA, J.N.,
tion of Canada, 317 p. KELLER, G.H., YANG, Z.S., AND FAN, Y.B., 1988, Marine dispersal and
WALKER, R.G., 1992, Facies, facies models and modern stratigraphic deposition of Yellow River silts by gravity-driven underflows:
concepts, in Walker, R.G., and James, N.P., eds., Facies Models: Re- Nature, v. 332, p. 629–632.
sponse to Sea Level Change: Geological Association of Canada, p. 1–14. YANG, C.S., AND SUN, J.S., 1988, Tidal sand ridges on the East China Sea
WALKER, R.G., AND HARMS, J.C., 1971, The “Catskill Delta”: A prograding shelf, in de Boer, P.L., van Gelder, A., and Nio, S.D., eds., Tide-
muddy shoreline in central Pennsylvania: Journal of Geology, v. 79, Influenced Sedimentary Environments and Facies: Dordrecht, The
Netherlands, D. Riedel Publishing Company, p. 23–38.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 293

A REEXAMINATION OF FACIES MODELS FOR CLASTIC SHORELINES

H. EDWARD CLIFTON
6 Cielo Vista Drive, Monterey, California 93940, U.S.A.
e-mail: eclifton@earthlink.net

ABSTRACT: Currently accepted depositional models of coastal facies derive mostly from studies of modern depositional systems combined
with interpretations of ancient successions. Two factors, however, can limit the efficacy of such facies models. First, Holocene coastal
settings differ significantly from those in which many, if not most, ancient coastal sediments accumulated. Second, input to a model from
the rock record commonly is specific to a particular tectonic and oceanographic setting (which may be poorly constrained) and is not fully
exportable to other settings. This paper explores how these factors impact our interpretive capability.
Many, if not most, ancient shoreface deposits accumulated under conditions of progradation, a process relatively uncommon among
the world’s present shorelines. Instead, many modern postglacial coasts experience rising sea level and reduced sedimentation, which
enhances barrier-island development and influences the shoreface-to-shelf bottom profile and sand–mud distributions. Ignoring these
differences promotes inaccuracy in our facies models.
Often overlooked in the application of coastal facies models are variations imposed by texture, energy level, and tectonism. Sedimentary
structures on fine-grained sandy coasts differ substantially from those on a coast underlain by coarse-grained sediment. Deposits on a high-
energy coast are unlike those in a low-energy setting. Differing degrees of accommodation influence the nature of the preserved succession.
Some of the problems inherent in current facies models can be obviated by considering them as end members within a continuum of models
that incorporate different energy regimes, textural characters, and preservational modes.

INTRODUCTION and applications of the models and the impact of these problems.
It concludes by suggesting an alternative approach that would
The basic model for open-coast clastic facies is not a complex allow a broader applicability of the models.
three-dimensional model, as with deep-sea fan facies or deltaic
facies, but rather a simple shallowing-upward facies succession A DURABLE MODEL
that is perpetuated in a sheet sand to its landward and seaward
pinchouts. The model follows the premise of Walther’s Law, The shallowing-up coastal classic facies succession in one or
whereby the vertical ordering of facies reflects the lateral arrange- another of its various manifestations (Figs. 1, 2, 3) is familiar to
ment of facies in a conformably prograding system (Middleton, virtually all students of shallow marine sandstones. The model
1973). In such a case, beach and shoreface (used here as the has persisted with little modification for nearly 30 years, as
relatively steep concave-up surface that lies between a beach depositional facies models for other systems, such as those for
foreshore and a shelf or basin platform) deposits prograde over deep-sea fans, evolved dramatically (and some continue to evolve,
adjacent shelf sediment. The resulting stratigraphic succession, as demonstrated in this collection of papers). The long-term
where complete, is an upward progression of shelf–shoreface– effectiveness of this model derives in large part from a combina-
foreshore–nonmarine facies (Fig. 1). Minor variations on the tion of simplicity and consistency in process, geometry, and
central theme exist in the various published iterations of the preservation on wave-dominated coasts.
model (Figs. 2, 3, 4). All share a common motif of upward Although a complex array of processes influence coastal
progression from bioturbated muddy sediment of the inner shelf settings, the effects of shoaling waves are primarily responsible
to mixed mud and storm sand transitional to the sandy shoreface, for shaping the nature of the clastic facies on open coasts, al-
which is dominated by storm structures in the lower part and by though tides and biogenic processes may be locally significant.
cross-bedded sand in the upper part. At the top lie flat-bedded Waves may exist as “seas”, driven by local winds, or as “swell”,
beach foreshore deposits that are overlain by backshore or other generated by distant storms. Swell tends to have longer period
nonmarine facies. More complex models have been proposed for and to influence the seabed to greater depths than do local sea
open-coast accumulations in settings where base-level fluctua- waves. “High-energy” coasts are likely to be dominated by swell.
tions impose patterns different from the laterally continuous “Low-energy” coasts receive smaller everyday waves but can
systematic facies progression. These include the “forced regres- experience very large waves during storms.
sion” models of Plint (1988, 1991) and Posamentier et al. (1992) Waves move sediment by two mechanisms: by inducing
(Fig. 5) and the “transgressive incised shoreface” of Walker and water motion as they pass, and, at the shoreline, by generating
Plint (1992) and Bergman and Walker (1999) (Fig. 6). In these sustained flow in the form of shore-parallel longshore currents
models, stratigraphic associations differ from the basic model and seaward-directed rip currents. The water motion induced by
and the upward facies progression may be incomplete. passing waves takes a consistent and predictable pattern that
Although the basic concept of shallowing-upward facies suc- relates to shoaling changes in the wave form itself. Waves passing
cession has been widely applied to ancient open-coast accumula- into shallow water change from a rounded, nearly sinusoidal
tions, the specific models that exist do not cover all situations. form to one of sharp-crested peaks separated by broad, flat
Some include assumptions that are demonstrably erroneous. troughs. As the waves approach the beach they become increas-
Much of the problem relates to limitations in the scope of studies ingly asymmetric and ultimately break when they encounter
on which the models are based. This paper reviews the origins water depths slightly (1.2–1.4 times) deeper than the wave height.

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 293–337.
294 H. EDWARD CLIFTON

PROGRADATION

NONMARINE

UPPER SHOREFACE upward-shallowing


facies succession
LOWER SHOREFACE

SHELF

FIG. 1.—Characteristic mode of accumulation of shoreline deposits through progradation (shown here with a slight rise of relative
sea level during progradation), whereby shallow-water facies build laterally over deeper-water counterparts, generating an
upward-shallowing succession or, if sandwiched between transgressive episodes, a parasequence (see Harms et al., 1982).

A
FORESHORE high
TIDE
low

UPPER
MIDDLE
FWWB LOWER
SHOREFACE
OFFSHORE

MUDDY SUBSTRATE SANDY SUBSTRATE

F W W B = FAIR–WEATHER WAVE BASE

B
LITHOLOGY ENVIRONMENT

COASTAL PLAIN
COAL
PLANAR LAMINATION BEACH FORESHORE
0m

BREAKER ZONE
CROSSBEDDING
(UPPER SHOREFACE)

SWALY CROSS-
STRATIFICATION SHOREFACE
10
HUMMOCKY CROSS-
LOWER SHOREFACE/
STRATIFICATION
INNER SHELF TRANSITION
BIOTURBATED
SANDY SILTSTONE MID-SHELF

FIG. 2.—A) Beach-to-offshore profile in facies model of Walker and Plint (1992). Fair-weather wave base at base of shoreface. B)
Shallowing-up facies succession in facies model of Walker and Plint (1992).
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 295

A HIGH-ENERGY SHOREFACE ZONATION

FORE-
UPPER SHORE

MIDDLE
LOWER
SHELF TRANSITION SHOREFACE

B
"HIGH ENERGY" "LOW ENERGY"

DUNE OR MARSH DUNE OR MARSH

FORESHORE FORESHORE

UPPER SHOREFACE
>2m
SHOREFACE
LOWER
SHOREFACE
< 25 m MIDDLE
SHOREFACE TRANSITION

LOWER SHELF
SHOREFACE

TRANSITION

SHELF

FIG. 3.—A) Beach-to-offshore profile (high-energy) in facies model of Galloway and Hobday (1996). Features a “transition zone”
between shoreface and shelf. B) Shallowing-up facies succession in facies model of Galloway and Hobday (1996). Features a
“transition zone” between shoreface and shelf in both high- and low-energy sequences.

The waves then pass through a breaker zone and end as swash All of this movement involves little or no mass transport of
and backwash on the beach foreshore. water. But as waves enter the breaker zone, water is carried
Water moved by the passing of a wave follows a circular orbit, forward by the waves. The result is a combination of setup and
forward under the crest of the wave and backward under the setdown, whereby the sea surface is elevated adjacent to the
trough, and the diameter of the circle diminishes with depth until beach and depressed just seaward of the breaker line (Fig. 10).
the movement becomes insignificant (“wave base”). As a wave These changes in elevation create a hydraulic head that serves as
enters shallow water, the circle is deformed into an ellipse and the the driving mechanism for longshore and rip currents, unidirec-
water movement just above the seabed is essentially horizontal tional flows that constitute nearshore circulation cells (Fig. 11).
(Fig. 7). As long as the wave form is nearly sinusoidal, the forward These simple processes encompass the significant forces that
and backward movement of the water is symmetrical in both drive sedimentation in the nearshore, and they have prevailed
duration and velocity. But as the wave crests become peaked, the since waves first came ashore early in the earth’s history.
velocity profile of the orbital currents changes: the forward Wave-dominated shorelines typically have a simple geom-
motion under the wave crests becomes stronger and of shorter etry. In plan view, the shorelines tend to be two-dimensional and
duration than the seaward motion (Fig. 8). uncomplicated. In profile, shoaling waves create a relatively
The velocity asymmetry of oscillatory (orbital) motion has steep, concave-up shoreface (Fig. 12) that extends seaward from
several geological ramifications. First, because bedload trans- the beach foreshore and merges offshore with a much flatter shelf
port is on the order of the cube of the velocity, the asymmetry or basin platform (Johnson, 1919). On prograding shorelines, the
has the capability to drive sediment on the bed in a landward shoreface is an equilibrium surface, probably reflecting a balance
direction. Second, bedforms created by this flow face in a of seaward sediment transport during storms and the landward
landward direction. Third, the stronger landward flow can transport by shoaling waves between storms (Niedoroda et al.,
overcome thresholds for movement of clasts that cannot be 1984). The sediment typically coarsens toward the upper part of
moved by the weaker seaward flow, and, where the grain-size the shoreface, the most energetic part of the system, where waves
range is sufficient, results in a preferential landward movement and longshore and rip currents typically shape the bed into
of the coarser clasts and textural sorting within the nearshore dunes. The shoreface geometry may be complicated by the pres-
zone (Fig. 9). ence of breaker bars, but the overall system is one of general
296 H. EDWARD CLIFTON

A
FORESHORE

MLW
FAIR-WEATHER WAVE BASE
SHOREFACE
OFFSHORE- STORM WAVE BASE
TRANSITION
OFFSHORE

MLW = Mean Low Water

B LITHOLOGY AND PROCESSES FACIES

SWASH ZONE FORESHORE

STORM EROSION, UPPER


FAIR-WEATHER SHOREFACE
DEPOSITION

STORM AND
FAIR-WEATHER SHOREFACE
DEPOSITION

STORM-DOMINATED OFFSHORE-
TRANSITION

FEW STORM EFFECTS OFFSHORE

FIG. 4.—A) Beach-to-offshore profile in facies model of Reading and Collinson (1996). Fair-weather wave base defines base of
shoreface. B) Shallowing-up facies succession in facies model of Reading and Collinson (1996).

Alluvial-plain and coastal-plain aggradation

A bayline Shoreline migration

B Alluvial-plain aggradation

bayline Shoreline migration

subaerial erosion
C
highstand shoreline
lowstand shoreline

submarine wave erosion

FIG. 5.—“Normal” and “forced” regressions. A) “Normal” regression under conditions of rising sea level. Shoreline deposits build
seaward and upward. B) “Normal” regression under stable sea level. Shoreline deposits build laterally. C) “Forced” regression
under conditions of falling sea level. Shoreline deposits become detached from their former position and can translate a
substantial distance into the basin. Subaerial erosion occurs landward of the new shoreline deposits, which, because of wave
erosion, display a sharp, erosional base over some portion of the previously deposited shelf facies. Modified from Posamentier
et al. (1992).
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 297

A
A 1. SEA LEVEL (SL) AT HIGHSTAND

A, F 2. RAPID SL
FALL
A, B
H G F
A
E
5. EROSION, D
CONSTANT SL
“CLIFF” C
SL 4. RAPID, STEPPED B
SL RISE
FWWB 3. SLOW SL RISE

WAVE EROSION IN SHOREFACE

B
POST-TRANSGRESSIVE SHELF MUD

TSE2
PRE-TRANSGRESSIVE
SHELF MUD
TSE1
SHOREFACE SAND BODY

FIG. 6.—A) “Stepped transgressions” following a major fall in sea level (2). Rapid rise in sea level (4) can preserve shoreface deposits
(E, D, formed during pauses in the transgression), which are “trapped” against a sea cliff, where they will be encased in offshore
mud as transgression progresses. Erosion surfaces are marked by transgressive lags. Modified from Walker and Plint (1992). B)
Template and log signatures for a shoreface deposit preserved after a stepped transgression (TSE1 = initial transgressive surface
of erosion which cuts a “cliff”, TSE2 = resumed transgressive surface of marine erosion). Shoreface sand body accumulates during
brief stillstand of sea level between the two phases of transgression. After Bergman and Walker (1999).

simplicity (Fig. 13). Finally, because wave processes tend to ORIGIN AND APPLICATION OF THE MODEL(S)
straighten progradational shorelines, the process of prograda-
tion tends to produce simple sheet sand deposits, with minimal Among the earliest depositional facies models proposed for
geometric complication. shallow marine sandbodies was that of the barrier island. Barriers
There is also simplicity and consistency in preservation. Shore- are particularly prominent on the U.S. Gulf Coast and East Coast
line successions typically develop where the coast progrades and on the southeastern coast of the North Sea, all sites of early
through the addition and accumulation of sediment (Fig. 1). The coastal sedimentologic studies (e.g., Shepard and Moore, 1955;
result is the basic depositional facies model, reflecting the shal- Van Straaten, 1959; Shepard et al., 1960; Bernard, et al., 1963;
lowing-upward succession of facies (typically coarsening-up- Reineck, 1963; Hoyt et al., 1964). Simultaneously, geologists began
ward as well). Accumulations of coastal deposits are likely to be to recognize that depositional facies could be very useful in the
preserved only within a subsiding basin. Fluctuations in eustasy, exploration for petroleum hydrocarbons. As a result, “linear clastic
sediment supply, or rates of subsidence typically generate alter- shorelines” became an early analog for open-coast deposits, and
nations of marine regression (progradation) and transgression. barriers became almost synonymous with shallow-marine sand
During marine transgression, the landward migration of the deposits. It was noted that barriers contained an upward-coarsen-
shoreface equilibrium profile (Fig. 14) removes much of the ing lithologic succession (Weimer, 1961; Bernard et al., 1963;
previously deposited material and produces a surface of ravine- Shelton, 1965, 1967; Berg and Davies, 1968), and for some time it
ment (transgressive surface of erosion). As a consequence, stacked was virtually assumed that all shallowing-up shallow-marine
sets of shoreline deposits tend to consist of a stacked set of sandstones originated in a barrier. Selley (1969), in his survey of
progradational parasequences separated by erosional surfaces depositional environments, focuses almost exclusively on barri-
formed during intervening transgressions (Fig. 15). Erosional ers, referring to nearshore deposits (even sheet sands) as “barrier
surfaces formed at the base of tidal inlet deposits during trans- beach” facies. In SEPM Special Publication 16, Recognition of
gression may complicate the pattern (Clifton, 2003). Ancient Sedimentary Environments (Rigby and Hamblin, 1972),
298 H. EDWARD CLIFTON

WAVE TRANSLATION BOTTOM OSCILLATORY CURRENTS

SYMMETRIC ASYMMETRIC

"wave base"

FIG. 8.—Schematic representation of relative velocity (arrow thick-


ness) and duration (arrow length) of oscillatory flow at the
FIG. 7.—Water particles moved by passing waves in deep water bottom beneath shoaling waves. Flow is symmetric under
follow a circular orbit that diminishes to zero at depth. In sinusoidal waves and asymmetric under sharp-crested waves.
shallower water, the particles follow an elliptical orbit that Flow is landward under wave crests and seaward under wave
flattens downward into a simple back-and-forth motion at the troughs. Stronger landward velocities (1) generate landward
sea floor, forward under the wave crest and backward under bed-load transport, (2) create landward-directed cross-bed-
the trough. The landward-most set of motions shown here is ding and ripple lamination, and (3) overcome threshold ve-
exaggerated for convenience. locities of clasts too large to be moved seaward, thereby
contributing to selective landward transport of coarser clasts
and overall textural sorting on the beach and shoreface.
a chapter is dedicated to criteria for recognizing ancient barrier
coastline (Dickinsen et al., 1972), and scant attention is given to
strand plains. As late as 1982, McCubbin, in a discussion of sandy showed the upward lithologic progression attributed to barrier-
coastal environments, discusses strand plains but devotes most island deposits (upward coarsening, upward transition from
of his text to barriers. In my experience in the petroleum industry marine to nonmarine environments), the lateral facies relations
in the 1990s, I found a surprising number of shallow-marine were inconsistent with a barrier interpretation. They could not,
sandstones interpreted as linear barriers, even where the geom- however, offer an alternative interpretation.
etry did not support a barrier-island interpretation. The description by Curray et al. (1969) of a prograding
Early depositional facies studies of shallow-marine sandbodies strandplain at Nayarit, Mexico, provided an alternative to the
encountered difficulties with the barrier interpretation. Harms et barrier-island model: a sheet sand with an internal shallowing-up
al. (1965), in a description of the Fox Hills Sandstone in Wyoming, facies succession. The facies models shown in Figures 1–3 are
could not accept an earlier interpretation of the unit as a barrier consistent with either prograding strand plains or barrier islands,
deposit (Weimer, 1961). They noted that although the sandstone where the primary differences lie in the geometries and facies

200

threshold velocity for moving a pebble with a diameter of 2.5 cm


Onshore
100

Bottom
orbital
velocity 0
(cm/s)

Offshore 100

threshold velocity for moving a pebble with a diameter of 2.5 cm

200
10 20
0
Time (seconds)

FIG. 9.—Selective shoreward transport of a 2.5 cm pebble under shoaling waves in which the landward oscillatory flow under the
wave crest has a greater velocity than the seaward flow under the trough. Dashed lines indicate threshold velocity, areas in red
indicate intervals of transport (adapted from Komar, 1976). Threshold velocity is derived from Komar and Miller (1973).
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 299

Breaker
line
Wave crests

Setdown Setup

Rip current Rip current


SWL
Breaker line Breaker line

Longshore currents Longshore currents

Beach Beach

FIG. 10.—Setup and setdown of the sea surface (departure from FIG. 11.—Nearshore circulation cells where wave incidence is
the still-water line (SWL)) owing to the landward transport of parallel to coast. Circulation consists of unidirectional rip and
water within the surf zone. Depression of the sea surface longshore currents generated by the hydraulic head caused
(setdown) just outside the breaker line and elevation of the sea by setup and setdown of the sea surface in and near the
surface adjacent to the beach provides the hydraulic head for breaker zone.
nearshore circulation cells.

based shoreface deposits resting abruptly on shelf mudstone, or


associations. Ryer (1977) noted the asymmetry of shoreface suc- isolated sandstone bodies encased in shelf mudstone or shale
cessions, in which the shallowing-up facies progression is typi- (Snedden and Bergman, 1999). Although these latter deposits
cally capped by an erosional surface. He attributed the erosion commonly share similar characteristics with shoreline succes-
and the accompanying abrupt return to facies deposited in deeper sions (upward coarsening and upward increase in depositional
water to a landward shift in shoreface profile during transgres- energy), they lack the sheet geometry and lateral facies relations
sion (Bruun, 1962). normally found in a progradational unit.
The prograding strand-plain model, which may have first The advent of sequence stratigraphy created a new awareness
been proposed by Harms et al. (1975) and was expanded upon by of the potential importance of sea-level change on facies associa-
Harms et al. (1982), has been applied successfully in many basins tions. The progradational-shoreline model is premised on con-
throughout the world. It predicts that prograding shoreline sand- stant or slightly rising relative sea level (Fig. 1). New models were
stones become progressively coarser and/or cleaner in an upsec- proposed involving a fall of relative sea level, or “forced regres-
tion direction, an attribute readily identifiable in well logs in areas sion” (Plint, 1988; Posamentier et al., 1992). In these models, a
where rock data are sparse or absent. relatively rapid fall in sea level induces a seaward “jump” of the
Despite the success of the prograding shoreline models, they shoreline deposits to a new, topographically lower position (Figs.
did not explain several categories of deposits, such as sharp- 5, 6, 16, 17). The models differ in their extent of wave erosion and

FORE-
SHORE
SHOREFACE

INNER
SHELF

FIG. 12.—The shoreface is best defined as a morphologic feature that attends nearly all clastic shorelines. It has a relatively steep
concave-up surface that extends seaward from the beach foreshore and merges with the much flatter shelf or basin platform.
Breaker bars, as shown, may cover the upper part of the shoreface here.
300 H. EDWARD CLIFTON

TRANSGRESSION

PROGRADATION
TRANSGRESSION

PROGRADATION

TRANSGRESSION

FIG. 13.—Two-dimensional wave-dominated open coast, South PROGRADATION


Carolina, U.S.A. Two sets of breaker bars are clearly visible: a
continuous inner bar, and an outer bar on which waves break TRANSGRESSION
more sporadically. Ridge-and-runnel systems occupy lower
beach foreshore, particularly in the distance. Although a
barrier presently occupies this coast, progradation would
produce a shallowing-upward sheet sand.
FIG. 15.—Typical pattern of stacked shoreline successions. Shal-
lowing-upward progradational parasequences meters to tens
nature of subsequent transgression. Plint (1988) and Snedden and of meters thick are separated by transgressive surfaces of
Bergman (1999) envision wave erosion by fair-weather waves erosion (ravinement surfaces). Falls in sea level prior to trans-
across the shelf platform in advance of the new shoreline position, gressions can produce erosional sequence boundaries and
producing an extensive erosional base to the advancing shoreface incised-valley-fill deposits at the top of the parasequences.
deposits (Figs. 16, 17). Posamentier et al. (1992) envision wave
erosion limited to the position of the newly established shoreline,
and a gradational shoreface–shelf transition as progradation The forced-regression models have been much applied since
ensues (Fig. 5). Posamentier et al. (1992) and Snedden and Bergman their inception (Posamentier and Chamberlain, 1993; Ainsworth
(1999) invoke a steady transgression that isolates the lowstand and Crowley, 1994, to cite a few). The Bergman and Walker
deposits, whereas Bergman and Walker (1988) and Walker and (1988) transgressive-incised-shoreface model for the Cardium
Plint (1992) call on a stepped transgression, featuring sporadic has been used to interpret other less well-documented shorefaces
stillstands of the sea. During the stillstands, shoreline deposits (Pattison and Walker, 1992; Walker and Wiseman, 1995; Le
can accumulate against a wave-cut sea cliff (Fig. 6), thereby Roux and Elgueta, 1997; Bergman, 1999; MacEachern et al.
forming narrow, shore-parallel linear sandbodies encased in 1999), including, somewhat controversially, the Shannon Sand-
marine shale as transgression proceeds. stone (Bergman and Walker, 1999). As will be discussed, how-
ever, not all sharp-based shoreface deposits are necessarily
produced by forced regression.
Ichnofacies models, based on associations of trace fossils,
TRANSGRESSION
have also been proposed for wave-dominated coastal succes-
SHOREFACE
LAGOON BARRIER INNER SHELF sions. Early students of these deposits recognized that traces
SL 2 could be useful in the interpretation of shallow-marine sand-
SL 1 stone (Weimer and Hoyt, 1964; Harms et al., 1965; Howard,
1966). The trace Ophiomorpha was thought initially to be indica-
RAVINEMENT SURFACE tive of a shallow-marine environment (Weimer and Hoyt, 1964;
Harms et al., 1965), but subsequent studies found it in a variety
LAGOONAL SEDIMENT BARRIER-ISLAND SAND of other marine environments, including those at bathyal depths
(Kern and Warme, 1974). The trace Macaronichnus occurs in
TIDAL-INLET FILL SHELF SAND AND MUD
many beach foreshore facies (Saunders and Pemberton. 1986;
MacEachern and Pemberton, 1992), but the trace is present in
other paralic environments as well, including upper shoreface
FIG. 14.—Process of marine transgression is commonly associated
(MacEachern and Pemberton 1992; Male, 1992), lower shoreface
with the landward migration of a barrier island or barrier spit.
(Clifton, 1981), and tidal flats and channels (Clifton and Thomp-
Landward translation of the shoreface profile as sea level rises
son, 1978). A problem with trace fossils is that, for most, we have
(here from SL 1 to SL 2) creates an erosional surface (surface
no knowledge of the physical or chemical factors that limit their
of ravinement or transgressive surface of erosion) cut into
distribution.
previously deposited sediment. Because deeper-water sedi-
Several attempts have been made to associate ichnologic trends
ments (here inner-shelf sand/mud) accumulate on the ravine-
to lithologic trends in nearshore sediment. Howard (1966) identi-
ment surface, it marks an abrupt upward change to deeper-
fied patterns in the trace assemblages that corresponded to the
water deposits.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 301

A
Swaly
cross-stratified sand
Fair-weather wave base

Hummocky
cross-stratified sand

Rippled and bioturbated sand and mud

B
Swaly
cross-stratified sand

Hummocky
cross-stratified sand

Rippled and bioturbated sand and mud

C “Normal”, gradationally based shoreface succession


Coastal plain
Coal/backshore
0 Beach foreshore
Upper shoreface
cross-bedded ss

Sharp-based shoreface succession


Shoreface
swaly cross- deposited during a forced regression
stratified ss

10 m
Shoreface-shelf
transition Relatively thin,
Hummocky cross- erosionally based
stratified ss shoreface sand
body

Mid-shelf Mudstone interclasts


bioturbated ss-mds gutter casts

Outer shelf
bioturbated mds

FIG. 16.—”Forced regression” as visualized by Plint (1988). A) Prograding shoreline in which swaly cross-stratification defines the
lower shoreface, hummocky cross-stratification defines the shoreface–shelf transition, and rippled and bioturbated sand and
mud typify the mid-shelf facies. B) Result of “forced regression”, where falling sea level induces mid-shelf wave erosion and
shoreface sands accumulate on the erosional surface. C) Vertical succession of a “forced regressive” deposit compared to a
“normal” progression.
302 H. EDWARD CLIFTON

A
Relatively slow sea-level fall

FWWB
Rippled sandstone and mudstone

(FWWB = fair-weather wave base)


B

Relatively rapid sea-level fall


thin shoreface sand

New erosional shoreface profile

Relatively stable sea level

Lowstand shoreface sandbody

D Relatively rapid sea-level rise

FWWB
TSE Transgressive mud blanket

Isolated sharp-based shelf sand body ("offshore bar")

FIG. 17.—Model of formation of isolated shallow marine sandbodies by relative sea-level change. A) Relatively slow sea-level fall;
shoreline moves seaward relatively slowly. Reduced accommodation results in a relatively thin shoreface sandbody. B) Faster
relative sea-level fall; shoreline moves seaward at an increasing rate. Rate of sea-level fall exceeds subsidence, accommodation
is reduced to nil, and the shelf becomes emergent. C) Relatively stable sea level; shoreline incises in seaward position. Erosion
plus subsidence create space into which a new shoreface sandbody can prograde. D) Relatively rapid sea-level rise; shoreline
shifts rapidly landward. Sandbody is isolated in shelf mud. After Snedden and Bergman (1999), following Plint’s (1988) model
(Fig. 16).

lithologic changes in shallowing-up Cretaceous sandstones in the subject to biological evolutionary trends and may therefore be
Book Cliffs, Utah, U.S.A. (Fig. 18), and Howard et al. (1972) and somewhat time-specific. Comparisons with modern analogs are
Howard and Reineck (1981) describe the distribution of physical also inherently difficult owing to the typically limited view of the
and biological structures in beach-to-offshore transects at Sapelo sub-sea-floor section on modern coasts. The studies of modern
Island, Georgia, and Port Hueneme, California, U.S.A. More re- biological structures by Howard et al. (1972) and Howard and
cently, Pemberton et al. (1992) followed up on Howard’s work in Reineck (1981), for example, provide little data that bear on the
the Book Cliffs with a more detailed analysis of the traces associ- model developed by MacEachern and Pemberton (1992).
ated with the lithologic succession (Fig. 19), and in 1992 MacEachern
and Pemberton (1992) proposed an ichnofacies model for Creta- THE HOLOCENE HERITAGE:
ceous shoreface successions in the western interior basin (Fig. 20). ANALOGS FROM A NON-ANALOGOUS WORLD
The ichnofacies model is a useful adjunct to models based on
texture and sedimentary structures, and may provide a basis for At present, a glacio-eustatic highstand exists, following a
subtle environmental interpretations not possible on the basis of rapid and large sea-level rise (that began about 17,000 years ago),
physical structures alone (MacEachern and Pemberton, 1992). a consequence of the melting of continental glaciers that devel-
Unlike models based on physical features, ichnofacies models are oped during the Wisconsin glaciation. Sea level continues to rise
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 303

Foreshore

Upper shoreface

Plug-
shaped
burrow

Lower shoreface

Plural Teichichnus
curving
tubes

Offshore

Ophiomorpha Asterosoma Snail trails

FIG. 18.—Distribution of traces in a shallowing-up succession in the Cretaceous Blackhawk Formation, Book Cliffs, Utah, U.S.A.
Constructed from data provided by Howard (1966).

on many coasts (albeit at a greatly reduced rate), and much of the which exists because the sediment supply is insufficient to fill the
world’s coastal area remains in a state of slow transgression. In basin landward of the barrier. Under this concept, the barrier
contrast, the coastal successions to which the classic shoreface continues to retreat to landward until a new progradational
model applies are, by definition, progradational. The attempts to episode begins. As noted by Suter and Clifton (1999), the biggest
incorporate observations from a still somewhat transgressive pitfall to using modern analogs is preservation potential. The
world into a progradational model have created a number of most common record of geologically preserved barrier islands
problems for our models of open-coast clastic facies. may be as the landward-most part of a progradational sand sheet
(Fig. 21).
Barriers
Sand–Mud Distribution
Presently, about 15% of the world’s coastline is fronted by
sandy barrier islands or barrier spits (Glaeser, 1978), much of it on The fact that many of the present-day coasts, particularly
U.S. coasts, and it is unsurprising that the linear sandbodies those with barriers, are sand-poor has also directed our thinking
represented therein provided an early, and widely used, explora- about the distribution of sand and mud in ancient coastal sys-
tion model. The origin of barrier islands has been much debated tems. Where sand is more or less confined to a barrier’s shoreface,
(see discussion in Davis, 1994), but most seem to form as a result a transition from sand to mud is likely to coincide with the base
of landward transport and upward accretion of sand (Davis, of the shoreface. Because so many of the early studies of coastal
1994). Many, if not most, modern barriers seem to be accumula- facies focused on barrier islands, it became generally accepted
tions of sand that are migrating landward as part of a slow that a transition from sand to mud defines the base of the
transgression (Kraft et al., 1973; Boyd, et al., 1992). A key factor in shoreface. This concept persists today in our models of prograd-
their development is probably a paucity of sand in the open-coast ing coastal deposits. In nearly all of them, the base of the shoreface
system. It is generally accepted that during a rise in relative sea coincides with a downward textural transition from sandstone to
level much sediment is trapped in rivers and estuaries and that shale (Figs. 2, 3, 4).
the amount of sediment, particularly sand, delivered to the open Under conditions of progradation, sand is likely to be far more
coast in many settings is greatly reduced. Waves mobilize the abundant and can extend far out onto the shelf. North of the
available sand and concentrate it along the coast into a barrier, mouth of the Columbia River, for example, the transition from
304 H. EDWARD CLIFTON

LITHOFACIES TRACE FOSSIL ASSEMBLAGES


FORESHORE 1 2 3 4 5 6 7
0

UPPER SHOREFACE

5
PROXIMAL MIDDLE SHOREFACE
DISTAL MIDDLE SHOREFACE

PROXIMAL LOWER SHOREFACE


10

DISTAL LOWER SHOREFACE


PROXIMAL LOWER SHOREFACE

15

DISTAL LOWER SHOREFACE

PROXIMAL OFFSHORE
20 m
TRACE FOSSILS
Helminthopsis A A O O - - -
Anconichnus C C A C - - -
Planolites A A C O - - -
Chondrites - O C C - - -
A = abundant Teichichnus - C C O - - -
C = common Thalassinoides - C C O - - -
O = occasional Terebellina S C C O - - -
S = sparse
Rosselia - O O O - - -
* = opportunistic following Rhizocoralium - S - - - - -
storm sedimentation
Skolithos S* S* O* C* C - -

Escape Burrow - S* O* A* S - S
Ophiomorpha - S* S* C* C O S
Arenicolites S* S* S* - S - -

Diplocraterion - - S* S* S - -

Palaeophycus - O* C* C* C O S

FIG. 19.—Trace-fossil distributions in core from the Spring Canyon Member of the Blackhawk Fm., Price, Utah, U.S.A. From
Pemberton et al. (1992).

sand to mud lies in a water depth of about 40 m near the river Shoreface Profile
mouth (Fig. 22). The transition extends to the north across the
shelf in progressively deeper water, ultimately reaching about 90 The relief of shoreface profiles in our present post-transgressive
m a few tens of kilometers north of the river mouth. Off north- world is variable (Fig. 25), depending on energy level and the
central California, the sand–mud transition lies consistently at a seafloor configuration prior to the transgression. The range in
water depth of about 60 m. (Fig. 23). Even off some barrier coasts, water depths at the base of modern shorefaces led Galloway and
such as that of New Jersey, sand extends offshore well onto the Hobday (1996) to conclude that that the thickness of shoreface
shelf (Fig. 24). Such sands may not be coeval with those on the facies successions spans some 2 to 25 meters (Fig. 3). This variation,
barrier but could be indistinguishable from them in the rock however, occurs on erosional coasts. Prograding coasts, which
record. Depositional facies models that equate the sandstone– provide the analog for nearly all ancient shoreline succession,
shale transition with the base of the shoreface may be valid for show less variability. On the prograding high-energy coast of
some deposits, but they do not provide an encompassing gener- southern Washington state, U.S.A., the break in slope that defines
alization. the shoreface–shelf transition occurs at a water depth of about 10
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 305

Backshore Psilonichnus
Ichnofacies
High Tide
Foreshore

Low Tide
Upper suspension
Shoreface feeding Skolithos
Ichnofacies

Middle
Shoreface

Lower
Shoreface
Minimum
wave base
Transition

deposit Cruziana
feeding Ichnofacies
Upper
Offshore

Lower
Offshore grazing &
foraging Maximum
wave base
Shelfal Zoophycos
Ichnofacies

FIG. 20.—Ichnofacies model for the shoreface, after Pemberton et al. (1992).

m (Fig. 26). The progradational coast of Nayarit, Mexico, a region ridges is generally coarser than that on their lee sides. Most
of somewhat lower wave energy, has the base of the shoreface at shoreface-attached ridges consist of an upward-coarsening ac-
around 6–7 m (Fig. 27). The base of the shoreface on Galveston cumulation of storm-event beds (Snedden et al., 1994;
Island, a prograding part of the Gulf of Mexico also with relatively Hoogendoorn and Dalrymple, 1986; Rine et al., 1991; Dalrymple
low wave energy (Morton, 1994), lies at a water depth of about 6 m and Hoogendoorn, 1997), although some show evidence of both
(Fig. 28). In contrast, the base of the shoreface off Padre Island, an storm and tidal influences in their internal structures (Antia et
eroding part of the Gulf coastline (Morton, 1994), lies in water al., 1994; van de Meene et al., 1996). Coring and/or high-
depths that approach 20 m (Fig. 25). On the basis of these observa- resolution seismic profiling show that many contemporary
tions it seems likely that, barring unusual rates of accommodation, ridges are compound features, composed of an upper part
the thickness of individual shoreface succession in the strati- shaped by modern processes and a core derived in an earlier
graphic record is not likely to exceed 10–12 m. setting (Rine et al., 1991; Snedden et al., 1994).
Many modern shorelines, particularly those on the U.S. Shoreface-attached ridges appear to be phenomena associ-
Atlantic coast, are fringed by shoreface-attached ridges, linear ated with a retreating shoreline (McBride and Moslow, 1991). As
bodies of sand, or sand ridges, that rise above the adjacent sea the shoreline shifts landward, some of the ridges are left behind
floor (Snedden et al., 1984; Hoogendoorn and Dalrymple, 1986; as isolated features on the inner shelf (Fig. 29), where they may be
Antia et al., 1994; van de Meene, et al., 1996; Dalrymple and further modified by shelf processes (Swift et al., 1986). Although
Hoogendoorn, 1997). The ridges can be tens of kilometers long, shoreface-attached ridges can be imposing coastal features, their
0.7 to 8 km wide, and 5 to 40 m high, and they are composed of association with transgressive coasts minimizes any importance
fine to coarse sand. The ridges typically lie oblique to the as part of a progadational shoreline model. Any preservation is
shoreline and tend to be asymmetric, with side slopes ranging most likely as isolated linear shelf sand bodies within a shelf
from < 1° to a maximum of 7°. Sediment on the stoss sides of succession (Swift and Parsons, 1999).
306 H. EDWARD CLIFTON

1. Sedimentation > base-level rise, progradation

2. Sedimentation < base-level rise, formation of barrier and lagoon

3. Sedimentation << base-level rise, transgression

4. Sedimentation > base-level rise, progradation

Shoreface deposits Bay fill


Tidal deposits and bay fill Inner-shelf deposits

FIG. 21. Schematic diagram showing the development, landward migration, and stranding of a barrier in response to changes in the
balance of sedimentation and relative sea-level fluctuation. During progradation in step 4, barrier maintains its identity until the
embayment created by the barrier is filled, at which time the coast converts to a strand plain.

TECTONIC SETTING AND GRAIN SIZE tion is a highly significant but commonly underrated parameter
in determining the nature of coastal facies.
Most of the early studies of modern open-coast systems were Grain size largely determines the slope of a beach–nearshore
conducted on tectonically passive margins of the U. S. Atlantic system. Coarse systems are steep, the waves break near the beach,
and Gulf coasts and the German and Dutch coasts of the North and wave energy tends to be reflected back into the ocean (Wright
Sea, where rivers with low gradients cross broad coastal plains et al., 1979). Fine beach–nearshore systems slope more gently,
and deliver fine sand to the shoreline. As a result, the emerging with the result that waves break farther from the shoreline and
models were premised on a nearshore system composed of dissipate their energy across a wide surf zone, in which bars and
uniformly fine-grained sand. These models were corroborated by troughs are likely to develop. It is commonly assumed that high-
studies in the Book Cliffs and elsewhere made on rocks of similar energy beaches are steeper than their lower-energy counterparts,
texture. The generalizations drawn in these studies, however, fail but studies of modern beaches have demonstrated just the re-
in varying degrees when applied to coarser-grained open-coast verse: they are more gently inclined (Komar, 1976). The assump-
deposits, particularly those in tectonically active settings. tion probably derives from erroneously mentally associating
Sedimentologists have tended to consider grain-size distribu- coarse beaches with high energy.
tions mostly to be reflective of processes of transport and depo- Grain size is a major influence in the size and shape of
sition, hence the numerous, largely unsuccessful, attempts to bedforms. This has long been known for unidirectional flow, but
reconstruct ancient depositional environments from textural pa- the textural relation may be even greater for wave-generated
rameters. While it is true that processes, in part, influence the structures. Figure 30 shows the spacing of symmetric ripples as a
texture of the sediment, sources and delivery systems also play a function of maximum bottom orbital velocity and grain size for a
significant role, particularly in the marine environment. Waves ten-second wave. Only very small ripples form in fine to very fine
can work only the sand population provided, and textural varia- sand at the same orbital velocities that generate megaripples in
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 307

200 m 100 m 50 m

MUD
VERY Willapa
FINE Bay
SAND

RELICT
SAND

Columbia R.
0 10 KM

FINE SAND

FIG. 22.—Distribution of sand and mud on the continental shelf off the mouth of the Columbia River, Pacific Northwest Coast of
the United States. After Nittrouer et al. (1986).

medium to coarse sand. It is unlikely that bedforms capable of


POINT ARENA
generating cross-bedding sets more than a few centimeters thick
can develop in fine to very fine sand under purely oscillatory
flow.
A comparison of texturally disparate coasts that have similar
wave climates illustrates the degree to which texture influences
facies character. Study of medium- to coarse-grained sandy
nonbarred nearshores in southern Oregon (Clifton et al., 1971)
100 showed that the sedimentary structures were arrayed in simple
shore-parallel patterns that reflected the transitions in the shoal-
ing waves (Fig. 31A). Asymmetric ripples dominate in the
150 offshore area, converting to decimeters-high lunate megaripples
SAND
in the area of most intense wave buildup just seaward from the
MUD surf zone. Landward migration of these bedforms produces
RELICT SEDIMENT
50 landward-dipping trough cross-bedding. Within the surf zone,
the bed is essentially flat; small, low-amplitude, transitory
500 ripples form between intervals of sheet flow as the waves
contours in meters
passed overhead. Adjacent to the beach foreshore, the bottom
0 50 km again becomes irregular at the interface between surf and swash
POINT REYES zones. Bedforms here faced seaward and produced seaward-
dipping cross-bedding. Within the foreshore, the bed is planar,
and the sediment contains gently inclined or planar parallel
FIG. 23.—Distribution of sand and mud on the shelf off central lamination.
California. The sand–mud transition lies at a water depth of The zones of sedimentary structures shift back and forth with
about 60 m, well seaward of the shoreface. After Drake and changes in wave climate and tides, producing assemblages of
Cacchione (1985). structures. An offshore–nearshore transition zone contains ripple
308 H. EDWARD CLIFTON

REHOBOTH BAY LAGOON, DELAWARE

0m

10

15

BARRIER SAND 0 3 km

SHALLOW MARINE SAND

LAGOONAL MUD, PEAT

PLEISTOCENE (After Kraft et al., 1973)

FIG. 24.—Distribution of sand on the shelf adjacent to the barrier at Rehoboth Bay, Delaware (arrow, above), after Kraft et al. (1973).
Shallow marine sand may be an older palimpsest deposit, but it demonstrates the lack of correspondence of the base of the
shoreface with the sand-mud transition.

A TRANSGRESSIVE SHOREFACE
CAPE ROMAINE, SOUTH CAROLINA

3–4 m
LAGOONAL SEDIMENT

TSE

B TRANSGRESSIVE SHOREFACE
SOUTH PADRE ISLAND, TEXAS
0m

10
BARRIER SAND A

LAGOONAL SAND & MUD 20 B


HOLOCENE DELTAIC DEPOSITS

LATE PLEISTOCENE DEPOSITS


0 5 km

FIG. 25.—Variability in transgressive shoreface profiles, Cape Romaine, South Carolina, U.S.A. (after Hayes and Sexton, 1989) and
South Padre Island, Texas, U.S.A. (after Morton, 1994).
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 309

LONG BEACH, WASHINGTON


A PROGRADING HIGH-ENERGY SHOREFACE

10

20 m SHOREFACE
SHELF

5 km 0

FIG. 26.—Shoreface profile on a high-energy, prograding coast, Long Beach, Washington, U.S.A. Break in slope defines the shoreface–
shelf boundary lies a water depth of about 10 m. After Dingler and Clifton (1994).

A B

PACIFIC
OCEAN
PROGRADING
NAYARIT, MEXICO SHORELINE
NAYARIT, MEXICO

C COSTA DE NAYARIT, MEXICO

10 km

Beach Ridges
Pacific Ocean

10 m

REGRESSIVE SAND
SILTY SAND
0 5 km SHELF MUD
MARSH, LAGOONAL FACIES

FIG. 27.—A) Location, B) plan view, and C) profile of the prograding coast at Nayarit, Mexico. Break in slope that defines shoreface–
shelf boundary lies in about 6-7 m of water. After Curray et al. (1969).
310 H. EDWARD CLIFTON

PROGRADATIONAL SHOREFACE
GALVESTON ISLAND, TEXAS
Galveston Island

0m

10

0 5 km
20

FIG. 28.—Profile across Galveston Island, Texas, U.S.A. Break in slope that defines shoreface–shelf boundary lies in about 6–7 m of
water. After Morton (1994).

B NORTH
EDISTO C
A RIVER

B 2m 4m 10 m
6m
A

C'
N

5 km
A' B'

C CHARLESTON, S. C., PROFLES


C C'

10 m

B B'

EBB-TIDAL
DELTA
10 m

A A'

10 m

FIG. 29.—Shoreface-attached ridges and isolated ridges on inner shelf, off the North Edisto River, Charleston, South Carolina, U.S.A.
A) Index map. B) Plan view. C) Cross section. Note break in slope at about 5 m on profiles A–A' and C–C', which is probably the
wave-cut shoreface on this complex coast.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 311

300 VELOCITY LIMIT — 10 SECOND WAVE

200
40
20 cm
FLAT BED
100
80
200 10 cm
60
50
40 RIPPLES
30 100
l = 5 cm
Um 20

(cm/s)
NO MOVEMENT
10

T = 10 SECONDS
2.00 1.00 0.50 0.250 0.125 0.062
GRAIN SIZE (mm)

FIG. 30.—Variation in ripple size (spacing = l) as a function of grain size and maximum bottom orbital velocity (Um) under a wave
with a period of 10 seconds. After Clifton (1976).

lamination, something similar to swaly cross-stratification, and tion. The process resulted in the finer sand moving offshore even
landward-facing trough cross-bedding generated under condi- as the coarser sand was being driven shoreward (Fig. 34). This
tions of heavy seas. A surf assemblage includes planar lamination process provides an effective means of textural segregation of
and landward- and seaward-dipping trough cross-bedding. A sand on the upper shoreface (Komar, 1976).
surf–swash transition assemblage contains planar lamination Texture can also influence processes in the nearshore. A
interrupted by wedges or troughs of seaward-dipping cross- striking example occurs on the Surinam coast of South America,
bedding. Progradation of a high-energy, non-barred, fair-weather, where suspended fine sediment discharged from the Amazon
coarse sandy, wave-dominated shoreline produces a stacking of River accumulates in the nearshore area. A zone of fluid mud
these assemblages in an upward-shallowing succession (Fig. concentrated near the shoreline damps about 95% of the wave
31B). In contrast, a fine sandy high-energy, nonbarred nearshore energy and transforms the incoming waves to a solitary wave
lacks the larger bedforms (Fig. 32A), and the vertical succession form (Wells and Coleman, 1978; Rine and Ginsburg, 1985).
shows a dominance of planar or gently undulating stratification Solitary waves are waves of translation that transport water
(Fig. 32B). (and mud) shoreward where the mud is trapped against the
Where the nearshore is composed of gravel, large straight- beach. Texture also controls rates of bioturbation, which tend to
crested, two-dimensional ripples or megaripples predominate be most rapid in fine to very fine sand and diminish as the grain
(Fig. 33A). Stratification typically is difficult to delineate in the size either increases to coarse sand or gravel or decreases to
gravel beds, and the resulting nearshore succession consists of mud.
well-segregated layers and lenses of apparently structureless
gravel interbedded with sand showing flat or inclined lamination EFFECTS OF BARS IN NEARSHORE SYSTEMS
and cross-bedding (Fig. 33B).
Sediment of different caliber can be transported in different Nearshore systems with breaker bars or other bars are inher-
directions under the same set of waves. The selective shoreward ently more difficult to study than are nonbarred coasts, particu-
transport of the larger clasts by asymmetric orbital currents larly on high-energy coasts with intense longshore and rip cur-
under shoaling waves has been noted (Fig. 9). Where mega- rents. As a result several detailed open-coast studies that are
ripples or other large bedforms exist, the asymmetric flow may applied to facies models were conducted on nonbarred coasts
result in coarser sand moving landward as part of the bed load, (Clifton et al. 1971; Howard and Reineck, 1981).
and finer sand moving seaward as part of a suspended load Bar–trough systems, however, are common on many, if not
(Inman and Bowen, 1963). While diving in the southern Oregon most, coastlines, and probably form part of nearly all ancient
surf zone, we noted that fountains of suspended sand commonly open-coast successions. The development of bars and troughs is
erupted from the lee sides of megaripples as the landward surge commonly linked to nearshore circulation cells of longshore and
of a wave diminished. The cloud of sediment would then drift rip currents (Fig. 11). Typically, on modern coasts, the location of
seaward under the offshore component of the orbital motion. offshore bars adjacent to a beach is readily seen from the breaking
Observations of both the clouds of sediment and dye streams pattern of waves in the bar crests. The studies noted below have
released as the sand fountains erupted indicated that the sus- focused on the facies of bar–trough systems on modern open
pended sediment never settled landward of the point of origina- coasts.
312 H. EDWARD CLIFTON

A
NEARSHORE
Buildup Breaker Swash
OFFSHORE zone zone zone

Asymmetric Lunate Outer Inner Inner


ripple megaripple planar rough Planar
facies facies facies facies facies

OFFSHORE NEARSHORE

B TEXTURE STRUCTURES FACIES


GRAVEL CS MS FS VFS
0m
Planar
lamination FORESHORE

Landward,
seaward
x-bedding; UPPER
Planar SHOREFACE
5 lamination

Landward,
x-bedding

Ripple, planar LOWER


lamination SHOREFACE
TO
10 Ripple
lamination; INNER
Bioturbation SHELF

FIG. 31.—A) Sedimentary structural facies in the non-barred nearshore (upper shoreface) in coarse sandy sediment on the high-energy
coast of southern Oregon, U.S.A., under fair-weather conditions. B) Vertical succession produced by progradation of such a
system.

Bars and troughs may be parallel to the shoreline and can lunate megaripples. In the longshore trough and rip channel,
occur in multiple sets of two or three, such as those on the Texas medium to small subaqueous dunes migrate in the direction of
Gulf Coast (Hill and Hunter, 1976). Here, breaking waves shape flow, respectively, producing longshore- and offshore-directed
bar crests into a plane bed; the sea floor in deeper water is covered cross-bedding (Fig. 36).
with wave ripples. Longshore currents stronger than about 0.5 Some coasts are characterized by irregular bar systems.
m/s generate small dunes that produce medium-scale, shore- Davidson-Arnott and Greenwood (1976) describe the facies that
parallel cross-bedding. Hill and Hunter (1976) note that intense form in mostly medium- to-fine-grained sand along the shore of
bioturbation destroys physical structures that lie 30 cm or more Kouchibouguac Bay, New Brunswick. Two sets of bars occur,
beneath the sediment–water interface. broadly shore-parallel, but with much irregularity. Bar crests
Where waves approach a coast obliquely, the bars and troughs here are composed of a combination of flat bedding and cross-
are likely to develop an en echelon pattern, in which individual bedding, and the troughs are largely underlain by ripple-lami-
bars are oblique and attached to the shoreline (Fig. 35). Study of nated sand. Rip channels that cut through the inner bar are
an attached oblique bar on the southern coast of Oregon (Hunter underlain by seaward-facing ripple lamination and cross-bed-
et al., 1979) showed that the sedimentary facies reflect the circu- ding.
lation cell. Water flows landward across the bar in the form of In all three examples, the bars shift landward and seaward as
very asymmetric oscillatory flow, generating either a flat bed or wave conditions change. The oblique bars on the Oregon coast
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 313

NEARSHORE
A Buildup Breaker Swash
zone zone zone

Asymmetric Outer Inner


ripple planar Planar
facies facies facies

OFFSHORE NEARSHORE

B TEXTURE STRUCTURES FACIES


GRAVEL CS MS FS VFS
0m
Planar FORESHORE
lamination

Planar
lamination UPPER
SHOREFACE
5
Planar or
undulatory
lamination

LOWER
Ripple SHOREFACE
10 lamination; TO
bioturbation
INNER
SHELF

FIG. 32.—A) Sedimentary structural facies in the non-barred nearshore (upper shoreface) in fine sandy sediment on a high-energy
coast under fair-weather conditions. Flatter beach–nearshore profile expands the surf zone relative to coarser shorelines. No
medium- to large-scale bedforms. Surf and swash zones are underlain by planar parallel lamination. B) Vertical succession
produced by progradation of such a system. Section lacks cross-bedding that typifies the upper shoreface of coarser shorelines.

also migrate laterally at rates of 100–200 m/ month and generate internal erosional surfaces marking the seaward migration of the
an envelope of bar–trough sedimentary facies (Fig. 37A). During troughs during progradation. The unidirectional currents can
progradation of this envelope, the currents in the trough land- generate bedforms in fine sand that would be shaped into a flat
ward from the bar erode previously deposited bar facies (Fig. bed or ripples by oscillatory flow (Fig. 39A). As a result, a
37B). Therefore the bar itself, although apparently the dominant progradational succession produced by a barred nearshore sys-
feature in the system, has a low potential for preservation. tem composed of fine sand can show abundant cross-bedding
Davidson-Arnott and Greenwood (1976) reach a similar conclu- that otherwise would be absent (Fig. 39B).
sion for the bars on the New Brunswick coast. Most interpreta-
tions of sand bodies as “offshore bars” are probably wrong. The FAIR-WEATHER OBSERVATIONS AND
vertical succession produced by a prograding nearshore bar– STORM-DOMINATED SYSTEMS
trough system contains an erosional surface that separates rip-
channel and longshore-trough facies from subjacent finer sand Most of our direct observations of nearshore processes
deposited on the seaward side of the bar (Fig. 38). come from studies conducted under conditions of fair weather,
The net effect of bar–trough systems is to enhance the unidi- when data can be collected most easily. Yet it is likely that
rectional flow of rip currents and longshore currents and to create processes operating during storms dominate much of the
314 H. EDWARD CLIFTON

A NEARSHORE Swash
Swash
OFFSHORE Build-up
Buildup Breaker
Breaker zone
zone
zone
zone zone
zone

Asymmetric Lunate 2-dimensional Inner


ripple megaripple megaripple Planar
facies facies facies facies

OFFSHORE NEARSHORE

B TEXTURE STRUCTURES FACIES


GRAVEL CS MS FS VFS
0m
Planar
lamination FORESHORE

Cross-bedding;
Planar
lamination;
gravel layers, UPPER
lenses SHOREFACE
5

Landward
cross-bedding

Planar or wavy LOWER


lamination SHOREFACE
TO
10
Ripple lamination
Bioturbation INNER
SHELF

FIG. 33.—A) Sedimentary structural facies in the non-barred nearshore (upper shoreface) in coarse gravelly sand on a high-energy
coast under fair-weather conditions. Large two-dimensional, straight-crested ripples occur in the gravel. These ripples tend to
face landward near the beach and be symmetrical in deeper water. B) Stratigraphic succession produced by progradation of such
a system.

nearshore stratigraphic record. Storms influence almost all transport become important (Fig. 41). The enhancement of rip
coastlines, and, compared to fair-weather waves, are capable currents and their extent into deeper water during storms (Fig.
of eroding, transporting, and depositing vast quantities of 42) provides a mechanism for transporting a wide range of grain
sediment. Analysis of wave records along most coasts indi- sizes to or beyond the base of the shoreface. Shoaling waves
cates a pattern where most of the time is occupied by fair- following the storm drive much of this material back onto the
weather conditions, a small but significant component of time upper shoreface, but some of the coarser grains (small pebbles,
is occupied by typical large annual storms, and a tiny fraction granules) are likely to be left behind, trapped in burrows or
of time is occupied by very infrequent major storms (Fig. 40). other depressions. The resulting bimodal sediment forms a
Each of these marks the sedimentary record in different ways, distinctive lower-shoreface facies in coarse sediment (Figs. 43,
depending on water depth. On the inner shelf, extreme events 44). In the absence of these pebbles, it may be very difficult to
are likely to produce the only physical structures in sediment distinguish between lower-shoreface and subjacent sandy-
otherwise dominated by bioturbation. The presence of sand in shelf facies. The upper shoreface is likely to be dominated by
this environment, however, by itself probably attests to trans- storm processes and consist of rip-current deposits, storm
port and deposition during storms of a wide range of sizes. On lags, and other storm-generated features (Fig. 43). Beach fore-
the shoreface, the effects of storm and fair-weather cross-shore shores are eroded during storms, and aggrade in fair-weather
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 315

A Onshore T1 T2 T3 T4 T5
200
100
Bottom
orbital
0
velocity
(cm/s) 100

200
Offshore 0 10 20
Time (seconds)

T1- Peak onshore flow. Suspended sand is concentrated in “roller” on lee side of megaripples

A B C

T2 -. End of onshore flow. Suspended sand fountains erupt from the lee sides of megaripples

A B C

T3 - Peak offshore flow. Suspended sand clouds from fountain move seaward.

C
B

T4 - End of offshore flow. Suspended sand rains out over megaripples to seaward.

B C

T5 - Next peak onshore flow. Sand deposited from the suspension clouds is caught in roller of
megaripple to seaward.

FIG. 34.—Differential transport of coarse and fine sand under the same set of waves. A) Velocity profile of currents generated at the
sea bed by a passing waves. B) T1–T5 correspond to times shown in Part A. Coarse sand moves shoreward as bed load, whereas
fine sand is thrown into suspension and drifts seaward, where it is trapped in rollers on the lee side of megaripples farther offshore.
Transport of sand to seaward in a field of megaripples under strongly asymmetric orbital flow.
316 H. EDWARD CLIFTON

wave
crest

Rip current Rip current

breakerline
breaker line
Bar Bar

Longshore currents Longshore currents


Beach

FIG. 35.—Nearshore circulation cells where wave incidence is oblique to coast. Longshore currents tend to flow in one direction only.
Such cells promote the development of attached oblique bars (shown in yellow).

-2

-2
-1

-1
RIP
CHANNEL BAR CREST

1 LONGSHORE TROUGH
2
3

BEACH FORESHORE
0 100 m
CONTOURS IN METERS AND HALF-METERS
A

-2

-2
-1

-1

1
2
3

0 100 m
CONTOURS IN METERS AND HALF-METERS
B

FIG. 36.—Bar and trough system on the southern coast of Oregon (Pistol River), U.S.A. A) Morphology and currents associated with
the system. B) Sedimentary structures associated with this system. Dunes occupy the longshore trough and rip channel, whereas
the bar crest is covered by lunate megaripples and/or a flat bed. Sediment in the longshore trough is coarser than that on the bar
or in the rip channel.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 317

A
FORESHORE DEPOSITS
MEAN SEA LEVEL

BAR DEPOSITS

TRANSITION
LONGSHORE- NEARSHORE-OFFSHORE
RIP- TRANSITION DEPOSITS
3m TROUGH DEPOSITS CHANNEL
DEPOSITS
100 m

B
PROGRADATION

MEAN SEA LEVEL

3m

100 m

FIG. 37.—Depositional facies in a prograding bar–trough system. A) Envelope of sedimentary facies generated by laterally migrating
shore-attached oblique bar system, as visualized in a cross section projected normal to the shoreline. B) Effects of progradation
of a laterally migrating shore-attached oblique bar system, as visualized in a cross section normal to the shoreline. Figure shows
the limited potential for preserving bar facies and the development of an internal erosional surface beneath the longshore trough
and rip channel. Coarse sediment from these environments sits abruptly over finer sediment deposited at similar water depths
outside the bar (nearshore–offshore transition). After Hunter et al. (1979).

intervals. Most beach foreshores thus consist primarily of fair- suggest that the gutter casts form when large waves drag gravel
weather deposits. back and forth on a sandy bed during a storm. They have not been
The dominance of storm effects on the upper shoreface im- seen under fair-weather conditions.
plies that features observed during fair-weather conditions may Another structure attributed by most workers to storms is
rarely be preserved. Examples include the lunate megaripples hummocky cross-stratification. This feature, which is common to
found in the high-energy nearshore of southern Oregon (Clifton nearly all shoreface and shelf fine-grained sandstone deposits,
et al., 1971). Decimeters-high dunes that migrate landward under has been observed forming only once in the natural environment
the asymmetric flow of shoaling waves, these features produce (Greenwood and Sherman, 1986). Many questions about the
medium-scale trough cross-bedding in the area just seaward of dynamics of its formation and its significance remain unan-
the breaker zone. Examination of many nearshore successions in swered.
the stratigraphic record shows that onshore trough cross-bed-
ding is uncommon (Fig. 45). The lunate megaripples are fair- BIAS TOWARD LOW-ENERGY COASTS
weather phenomena that are obliterated by the enhanced rip
currents and longshore flow that accompanies storms. Unsurprisingly, most of our knowledge about modern
Conversely, several structures that are common in the sedi- shoreface facies comes from the study of coasts with low wave
mentary record have never been seen during their formation. energy. The few studies of high-energy systems, where the
Gravel-filled gutter casts are common in pebbly nearshore sand- wave heights are routinely in the range of 1–2 m, have been
stones. These shore-normal structures commonly have steep, or focused on nearshore areas close to the shoreline. As a result,
even undercut, sides indicating nearly simultaneous cutting and our understanding of modern open coasts is strongly biased
filling (Chiocci and Clifton, 1991). Their orientation and shape toward low-energy systems. The result has been some errone-
318 H. EDWARD CLIFTON

TEXTURE STRUCTURES FACIES


GRAVEL CS MS FS VFS
0m
Planar FORESHORE
Lamination

Longshore or
seaward UPPER
crossbedding; SHOREFACE
5 Inclined planar
lamination

Parallel or wavy
lamination; LOWER
SHOREFACE
Ripple TO
10 lamination; INNER
Bioturbation SHELF

FIG. 38.—Vertical sequence produced by progradation of an attached, oblique bar system in a high-energy, coarse sandy setting.

ous generalizations that pervade the models of the wave-domi- the wave length (Sverdrup et al., 1942), and many geologists have
nated coastal facies. followed suit (Walker and Plint, 1992; Reading, 1996). This defi-
nition, however, does not consider the effect of wave height; large
Wave Base waves disturb the bottom at depth uninfluenced by smaller
waves of the same period or wavelength. The concept that wave
Perhaps the most broadly held misconception in the interpre- base equates with the water depth in which sediment first begins
tation of shoreface systems is that of the role of fair-weather wave to move implies that wave base is partly dependent on sediment
base relative to facies distributions. Reineck and Singh, in their caliber; under the same set of waves, a fine bed might be above
justifiably influential book on depositional sedimentary environ- wave base whereas a coarser bed might not be. Moreover, wave
ments (1973), state that the seaward limit of the shoreface corre- theory predicts that a bed of fine sand is mobilized by fair-
sponds to wave base, which in they identify as the “average weather long-period swell at water depths much beyond the base
maximum wave base”. From their studies in the low-energy Gulf of the shoreface (Fig. 46A). Shorter-period waves, such as those
of Gaeta, they concluded that the boundary between the upper characteristic of a low-energy coast, move fine sand sediment in
and lower shoreface corresponded to fair-weather wave base. In considerably shallow water (Fig. 46B). A coincidence of fair-
the low-energy setting of Long Island, New York, U.S.A., Shipp weather wave base and the base of the shoreface, however, is
(1984) found that the maximum depth to which fair-weather likely to be just that: a coincidence depending on wave height and
waves moved sediment corresponded with the base of the period and sediment grain size. It is noteworthy that for both
shoreface. Such studies were incorporated into the models for a long-period (10 s) and shorter-period (5 s) waves, a water depth
wave-dominated coast. Fair-weather wave base coincides with equivalent to one-half of the deepwater wave length lies well
the base of the shoreface in the models provided by Walker and seaward of the shoreface base on a prograding shoreline (Fig.
Plint (1992) and Reading (1996). The same relation is implicit in 46A, B).
the text of Galloway and Hobday (1996), who note that the lower Finally, geologists can interpret wave base in the stratigraphic
shoreface is influenced by both storm and fair-weather waves, record only by inference, such as the balance between physical
whereas fair-weather waves (other than long-period swell) have depositional structures and biogenic structures or the presence of
little effect on the shelf. Consequently, a number of workers mud layers in the section. In the first case, fine sand is commonly
postulate that fair-weather wave base defines the shelf–shoreface completely bioturbated even at water depths where everyday
boundary (e.g., MacEachearn and Pemberton, 1992; Maejima, waves ripple the surface. The ripples would indicate deposition
1993; Hettinger et al., 1994; Hart and Plint, 1995; Hampson and above fair-weather wave base, but the bioturbation could suggest
Storms, 2003). deposition possibly below storm wave base. Layers of mud can
Although this interpretation may be valid for some succes- accumulate in shallow water from the rapid settling of large
sions, it is invalid as a generalization. First, wave base is so volumes of silt and/or clay resuspended by storm waves or
variously defined that it has lost much of its currency. Most introduced by floods. Their presence is unrelated to fair-weather
geologists identify wave base as the greatest water depth in conditions. The extension of generalizations regarding wave
which passing waves disturb the bed, although some have used base, drawn from studies of modern low-energy coasts, is largely
the term to separate the zones of “normal” wave erosion and unwarranted.
wave deposition (Kowalewsky, 1982), which is postulated to
occur at a depth of about 10 m (Schwartz, 1982). Plint (1988) seems Wave Energy and Facies
to use this definition in his model of forced regression. Physical
oceanographers have placed wave base at a depth where waves Several workers have attempted to compare the facies of high-
begin to “feel bottom”, approximately equivalent to one-half of energy and low-energy coasts. Clifton (1976) contrasted the fair-
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 319

A NEARSHORE
Swash
OFFSHORE Buildup Breaker Reforming zone
zone zone zone

Asymmetric Plane-bed Bar-trough Inner


ripple bar surface facies planar
facies (shore-parallel facies
cross-bedding)

OFFSHORE NEARSHORE

B TEXTURE STRUCTURES FACIES


GRAVEL CS MS FS VFS
0m
Planar
lamination FORESHORE

Longshore or
seaward
crossbedding; UPPER
Inclined planar SHOREFACE
lamination
5

Parallel or
ripple lamination; LOWER
Bioturbation SHOREFACE
TO

10 Ripple INNER
lamination; SHELF
Bioturbation

FIG. 39.—A) Sedimentary structural facies in a barred high- to moderate-energy fine-grained nearshore under fair-weather
conditions. Longshore flow in the trough of the shore-parallel bar is strong enough to create small dunes that migrate alongshore
in the trough. B) Vertical succession produced by progradation of the system shown in Part A. Fine sand shows much shore-
parallel trough cross-bedding that would not be formed without unidrectional flow within a bar–tough system.

weather nearshore facies in sand of similar texture on the very low-energy coast in Georgia, U.S.A. Their analysis is based on
high-energy coast of southern Oregon, the moderately low- both box-core and vibracore analyses, which incorporates the
energy coast of southeastern Spain, and the very low-wave- effects of coastal storms. Their comparison shows a similar thin-
energy beach within Willapa Bay, Washington, U.S.A., (Fig. 47A). ning of the facies progression and loss of higher-energy features
The facies distribution becomes increasingly compressed as fa- in the low-energy setting (Fig. 48). The thickness of the shoreface
cies requiring relatively high velocities progressively disappear section that would result from progradation of these coasts
as wave energy is decreased. The upper-flow-regime plane bed differs from about 9 m in California to 2 m in the example from
that characterizes the inner surf zone of southern Oregon is absent Georgia. Cross-bedding, common in the shoreface deposits in
on the Spanish coast, and the lunate megaripples and cross-ripples California, is absent in the Georgia succession. Some textural
observed in Oregon and Spain are absent in Willapa Bay. As the differences may exist in these two examples. The California
wave energy of the setting diminishes, the progradational succes- nearshore is typified by fine- to medium-grained sand (Howard
sions (Fig. 47B) become thinner and increasingly impoverished in and Reineck, 1981), whereas the Georgia nearshore seems to be
sedimentary structures requiring relatively strong currents. composed of uniformly fine-grained sand (Howard et al., 1972).
Howard and Reineck (1981) contrast the facies succession As noted below, textural differences may outweigh variations in
from a high-energy coast in Southern California with that of a wave energy in shaping coastal facies.
320 H. EDWARD CLIFTON

100 comparison with a third shoreline, on the southeastern coast of


FAIR WEATHER Spain, helps to resolve the relative influences of texture and
PERCENT ambient energy. The environmental setting of this Spanish coast
OF
TIME
TYPICAL LARGE WINTER STORM is very similar to that of the Texas Gulf coast (fair-weather wave
heights in the range of 0.2–0.5 m, tidal range less than a meter),
EXTREME EVENT but texturally this tectonically active coast resembles that of
50 Oregon. Profiles (Fig. 50) show development of a succession of
structures similar to that seen on the Oregon coast. Even where
bars composed of fine sand lie off the beaches, the intervening
RECURRENCE INTERVAL (YR) troughs are occupied by gravel shaped into large 2-D ripples
2 5 10 20 50 100 like those shown in Figure 50B.
A comparison of outcrops representing each of these three
0 environments (Figs. 51–54) likewise shows greater similarity
0 2 4 6 8 10 12 14 16 18 between the Spanish and Oregon deposits. The primary differ-
WAVE HEIGHT (M), GALVESTON, TEXAS
ence between the two coarse-grained deposits is that the low-
energy succession is significantly thinner that that formed un-
der high-energy conditions, as a consequence of the deeper
FIG. 40.—Energy (wave height) frequency for Galveston Island, extent of wave influence on the high-energy California coast.
Texas, U.S.A. During most of the time small waves prevail Fine-grained, shallowing-up coastal succesions occur in the
(fair-weather conditions). A small, but significant, amount of Eocene Jackson Group of West Texas, where presumably they
time is occupied by large-winter-storm conditions, and very accumulated under conditions similar to those on the present
large storm (hurricane) waves occur very infrequently. This Texas Gulf Coast. Once exposed in now-covered uranium pits
pattern is typical for most coasts, although the wave heights in West Texas, these successions are thinner than might be
involved may differ (for example, on the central California expected on a fine sandy coast with greater wave energy, and
coast, waves 2 m high are fairly common, and typical large they also show more bioturbation than might occur on a high-
winter storm waves range from 4–5 m high). energy coast. The successions differ strikingly from those formed
in coarse sediment under similar oceanographic conditions off
Intuitively, it would seem that variations in energy regime the coast of Spain. Although ambient wave energy is a factor in
would constitute a major influence on the character of shoreline facies development, any interpretation of energy level must
facies. In reality, that influence can be difficult to resolve, largely take into account the textural factor.
because of textural complications. The nearshore profile off Low wave energy may be an important factor in shaping one
Padre Island (Fig. 49), where fine to very fine sand accumulates particular type of shallow marine deposit. The shoreface is gen-
in a low-energy setting (fair-weather waves 0.2–0.5 m high tidal erally presumed to be an equilibrium profile for a given set of
range > 1 m) differs markedly from that of the Oregon coast, wave conditions. Where equilibrium is not achieved, owing to an
(Fig. 31), where medium to coarse, pebbly sand accumulates in inability of the waves to redistribute the introduced coarse sedi-
a high-energy setting (fair-weather waves 1–2 m high; tidal ment, and/or insufficient time to reshape the profile as new
range 2–3 m). Much of the difference, however, may be due to sediment accumulates, and/or a very steep offshore gradient, a
the diverse textural character of these two environments. A Gilbert delta rather than a shoreface is likely to develop (Corner
et al., 1990; Postma, 1990).
Gilbert deltas are characterized by steeply inclined foresets in
STORM LAG tabular sets that can be tens of meters thick (Colella, 1988a, 1988b;
COARSE GRAVEL Nemec, 1990). Sediment transport down the face of the delta
occurs primarily by mass transport (Postma, 1984; Postma et al.,
POST-STORM LAG 1988). Many, if not most, Gilbert deltas are conglomeratic, finer
FINE GRAVEL sediment being more easily shaped into a shoreface. But in areas
of powerful waves, as along much of the U.S. West Coast, even the
coarsest gravel can be reworked into an equilibrium profile and
COARSE SAND
FAIR-WEATHER Gilbert deltas do not develop.
STORM CONDITIONS
CONDITIONS
MEDIUM SAND
SHARP-BASED SHOREFACE DEPOSITS

The models for coastal deposits that formed during a forced


FINE SAND regression derive almost entirely from the stratigraphic record.
EQUILIBRIUM SIZE Although Pleistocene deposits on the outer part of modern con-
tinental shelves are cited as contemporary examples (Posamentier
OFFSHORE 0 ONSHORE et al., 1992), little is known about the lithologic details of these
RATE OF SEDIMENT TRANSPORT deposits or the processes that attended their formation. The
concept that forced regression can produce an extensive ero-
FIG. 41.—Idealized diagram showing rate and direction of trans- sional base to shoreface deposits, as Plint (1988) postulated for the
port on the shoreface as a function of grain size under storm Cardium Formation, has been widely applied. Posamentier et al.
and fair-weather conditions. Sediment carried seaward by (1992) note that, although coastal deposits in the Viking Forma-
storm rip currents is reworked in the aftermath of the storm. tion in Joarcam Field, Alberta, Canada, seem produced by forced
On the lower shoreface, fine sand in equilibrium with the fair- regression, shorefaces are sharp-based only in their most proxi-
weather waves coexists with fine gravel deposited during the mal position, rather than over the entire width of their occur-
storm and left behind as a post-storm lag. rence, as in the Cardium. The difference between the two may
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 321

A
WAVE CRESTS

RIP CURRENT
BREAKER LINE

LONGSHORE CURRENTS

BEACH

WAVE CRESTS
B RIP CURRENT
BREAKER LINE

LONGSHORE CURRENTS

BEACH

FIG. 42.—Nearshore circulation cells as a function of wave energy. A) Small waves break in shallower water than do large waves B).
Larger waves also generate a greater degree of setup/setdown and thus have more strongly developed rip and longshore
currents. Rip currents under large waves extend farther offshore, probably to the base of the shoreface or slightly beyond.

reflect the finer grain size of the Viking sand or lower wave by oblique shore-attached bars, rather than from forced regres-
energy (Posamentier et al., 1992), or possibly a somewhat lower sion (Figs. 55, 56). In these deposits, the sharp contact separates
basinal gradient for the Cardium. coarser, cross-bedded sandstone from subjacent finer and more
Although some sharp-based shoreface deposits unquestion- bioturbated sandstone. The common presence of small pebbles in
ably reflect a forced regression, as for example where the ero- the sandstone below the contact and occasional intertonguing of
sional surface is of regional extent (Plint, 1988; Hadley and Elliott, the facies suggests that the erosional surface does not represent a
1993), other explanations exist. Figure 6, for example, shows major break in facies succession.
sharp-based shorefaces that result from incision during a stepped Parasequences in the Blackhawk Formation in the Book Cliffs,
transgression (Walker and Plint, 1992; Bergman and Walker, Utah, U.S.A., also contain sharp-based nearshore sandstones that
1988, 1999). As noted in a previous section, a prograding bar– may be unrelated to forced regression. Two types of contacts
tough system can create an erosional surface at the base of upper- occur (Fig. 57). One lies at the base of the upper shoreface
shoreface deposits (Fig. 38). Examples of this feature occur re- deposits, as noted by Howard (1972), where clean, medium-
peatedly in middle Miocene shoreline deposits in the Caliente grained cross-bedded sandstone sharply overlies finer biotur-
Range of California, U.S.A. (Clifton, 1981). Paleocurrent mea- bated sandstone of the lower shoreface. As with the Caliente
surements support the conclusion that the sharp-based upper- Range example, the break in succession is relatively minor and is
shoreface deposits here result from progradation of a coast marked probably attributable to a prograding bar–trough system. The
322 H. EDWARD CLIFTON

A STORM EFFECTS, HIGH-ENERGY, NONBARRED COARSE SAND


TEXTURE STRUCTURES FACIES
GRAVEL CS MS FS VFS
0m
Planar BEACH
lamination FORESHORE

Seaward
cross-bedding;
pebble and
cobble layers
5 UPPER
SHOREFACE
Seaward
cross-bedding;
Layers and
lenses of fine
pebbles

10
Scattered fine LOWER
pebbles; SHOREFACE
Bioturbation

Bioturbation INNER
HCS SHELF

B STORM EFFECTS, HIGH-ENERGY, NON-BARRED FINE SAND


TEXTURE STRUCTURES FACIES
GRAVEL CS MS FS VFS
0m
Planar BEACH
lamination FORESHORE

Planar
lamination

Planar or UPPER
5
undulatory SHOREFACE
lamination
Swaly cross-
stratification

LOWER
Hummocky SHOREFACE
10
cross-strata TO
Bioturbation
INNER SHELF

FIG. 43.—A) Vertical succession produced by progradation of a high-energy, nonbarred, coarse sandy nearshore, in which storm
effects predominate. Scattered pebbles on lower shoreface interpreted as reworked post-storm lags. Scattered fine pebbles in the
lower shoreface are interpreted as remnants of post-storm lags. B) Vertical sequence produced by progradation of a storm-
dominated, nonbarred, high-energy, fine-sandy nearshore. Cross-bedding may result from increased nearshore circulation flow.

second contact lies at the base of amalgamated storm sands. In the CONCLUSIONS
section below, muddy intervals separate the storm sets. Although
striking, the contact at the base of the amalgamated sandstone is The basic facies model for open-coast clastic deposits is a
probably comparable to that at the base of each of the subjacent simple upward-shallowing succession in a sand body bounded
storm sands. The absence of shale in the overlying section, to seaward by shelf deposits and to landward by nonmarine
because of either greater storm erosion in the shallower water or facies. The sand body may be linear, as in a barrier, or sheet-like,
inability of mud to accumulate in the more energetic environ- as in a strand plain. The model exists with minor variations in the
ment, defines the break. Such contacts cannot be traced laterally standard texts and has received broad application. It is flawed to
and are unlikely to represent forced regressions. a degree, in that it is based on a fairly limited set of modern
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 323

A 100
WATER DEPTH HIGH-ENERGY COAST
(m) 90
T = 10 S
80
1/2 L
70
FAIR
60 WEATHER
STORM
50

40

30

20
SHOREFACE
BASE THRESHOLD FOR MOVING FINE SAND
10
um = 0.17 m/s
0

0 1 2 3 4 5 6 7 8
FIG. 44.—Photograph of transition from upper shoreface to lower WAVE HEIGHT (m)
shoreface (just below10 cm scale) in the Plio-Pleistocene Merced
Formation, San Francisco, California, U.S.A. Upper shoreface B 30
has cross-bedded sand and gravel. Lower shoreface is marked WATER DEPTH LOW-ENERGY COAST
by a couple of meters of fine sand that bears scattered small (m)
T=5S
pebbles in stringers and isolated clusters.

analogs. Many studies of modern coasts have been made in areas 20


1/2 L
undergoing marine transgression, and conclusions drawn from
STORM
these coasts are skewed in their view of sandbody geometry,
shoreface profile, and sand–mud distribution. In addition, stud- FW
ies of modern coasts are biased toward fair-weather conditions
10
0° SHOREFACE
10% BASE
330° 30°
8 THRESHOLD FOR MOVING FINE SAND
um = 0.15 m/s
SHORELINE 6
0
TREND
300° 4 60° 0 1 2 3 4 5 6

WAVE HEIGHT (m)

FIG. 46.—Combination of wave heights and water depths in


Land which fine sand (D = 0.125 mm) will be moved by passing
270° n = 149 90° waves. A) 10-second waves. Combination of wave heights
and water depths in which fine sand (D = 0.125 mm) will be
moved by passing waves. B) 5-second waves. In both cases
movement occurs in water depths well seaward of the base of
the shoreface on a prograding coast. Water depths equal to
one-half the deep water wave length are deeper still.
240° 120°

and settings of low wave energy. Most modern studies have been
made on coasts in tectonically passive settings, where fine sand
predominates on the beach and shoreface. The influence of tex-
210° 150°
ture, as an independent variable, has largely been overlooked.
180° Many of the limitations on the basic model could be obviated
by considering the variations to be end members in a flexible or
FIG. 45.— Summary of 149 cross-bedding measurements in near- even multi-dimensional model based on parameters such as
shore facies of the Plio-Pleistocene Merced Formation. Red texture, sand supply, ambient wave energy, storm influence,
bars indicate orientation of gravel-filled gutter casts that are coastal morphology, and nature of base-level change (Fig. 58).
approximately normal to the shoreline. Most of the cross- Using this approach, the basic model as presented in most texts
bedding indicates south-flowing longshore currents. Very becomes specifically a model for a storm-dominated, moderate-
little cross-bedding is directed landward, indicating that lu- energy to low-energy setting with a moderate gradient in which
nate megaripples produced by fair-weather waves are rarely fine sand was in somewhat limited supply and base level was
preserved. After Chiocci and Clifton (1991). static (Fig. 59). Coastal successions in Pleistocene deposits on the
324 H. EDWARD CLIFTON

A LANDWARD
SYMMETRIC ASYMMETRIC
BIOTURBATED 2-D 3-D CROSS LUNATE FLAT
SEAFLOOR RIPPLES RIPPLES RIPPLES MEGARIPPLES BED

Oregon Coast
water depth = 1–30 m, wave height 1–3 m

Southeastern Spain
water depth = 1–5 m, wave height 0.2–0.5 m

Willapa Bay, Washington


water depth = 0.5–1 m, wave height < 0.2–0.5 m

B “LOW ENERGY”
“HIGH ENERGY”

“VERY LOW ENERGY”

FORESHORE FORESHORE

MLW MLW

5m 5 5

10 m 10

MLW = Mean low water


15 m

FIG. 47.— A) Distribution of sea-floor facies observed under fair-weather conditions on a high-energy coast (southern Oregon), U.S.A,
a low-energy coast (southeastern Spain), and a very low-energy coast (Willapa Bay Washington, U.S.A.), and B) comparison of
the hypothetical beach-to-offshore successions produced by progradation of facies shown in Part A.

California coast (Fig. 60) reflect accumulation in a storm-domi- Facies models have been proposed for open-coast sediment
nated, high-energy, barred setting with an abundant supply of in settings other than progradation at constant or slowly rising
sand, including coarse sand and gravel, under conditions of static relative sea level. In particular, models based on “forced re-
base level. Pleistocene successions on the southeastern coast of gression” have provided an alternative for explaining isolated
Spain (Fig. 61) were deposited in a similar setting, but under shallow marine sand bodies with sharp bases. The vertical
conditions of much lower wave energy. Successions like those succession produced thereby fits into the model as forming in
formed on the Texas Gulf Coast (Fig. 62) represent deposition in a storm-dominated, moderate-energy to low-energy setting
a similarly low-energy setting, but one dominated by fine sand, with a low gradient in which fine sand was in somewhat
in which fair-weather processes predominate in the preserved limited supply and falling (or fallen) base level (Fig. 64). Not all
deposit. Gilbert-delta deposits (Fig. 63) can be accommodated sharp-based shoreface successions, however, require sea-level
into the model as accumulating in a steep-gradient setting of low change. Prograding bar–trough systems and the simple amal-
to very low energy. Although the supply of gravel may be gamation of storm sands in a setting with limited sand supply
substantial, not enough sand enters the system to develop an can also produce erosionally based coastal sandstone deposits
offshore profile in equilibrium with the waves. (Fig. 65).
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 325

“HIGH ENERGY”
“LOW ENERGY”

FORESHORE
MLW MLW
SHOREFACE

5 5

TRANSITION
10 ZONE 10

15 15 m
OFFSHORE

20 m
MLW = Mean low water

FIG. 48.—A comparison of high- and low-energy beach-to-offshore sequences, Ventura to Port Hueneme, California, and Sapelo
Island, Georgia, U.S.A. After Howard and Reineck (1981).

ACKNOWLEDGMENTS BERGMAN, K.M., 1999, Cretaceous Sussex Sandstone in House Creek Field
(Wyoming, USA): Transgressive incised shoreface deposits, in
I am greatly indebted to Sandra Phillips, David James, and Bergman, K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies:
William Morris for their insightful reviews. They caught far too Sequence Stratigraphic Analysis and Sedimentologic Interpretation:
many typos and raised many salient points, and the paper has SEPM, Special Publication 64, p. 297–319.
benefited immensely from their comments. I also thank the BERGMAN, K.M., AND WALKER, R.G., 1988, Formation of Cardium Erosion
editors of this volume for providing a format for presenting this surfaces E5, and associated deposition of conglomerate, Carrot Creek
material. Field, Cretaceous Western Interior Seaway, Alberta, in James, D.P.,
and Leckie, D.A., eds., Sequences, Stratigraphy, Sedimentology; Sur-
REFERENCES face and Subsurface: Canadian Society of Petroleum Geologists,
Memoir 15, p. 15–24.
AINSWORTH, R.B., AND CROWLEY, S.F., 1994, Wave-dominated nearshore BERGMAN, K.M., AND WALKER, R.G., 1999, Campanian Shannon Sandstone:
sedimentation and ‘forced’ regression: post-abandonment facies, great an example of a falling stage systems tract deposit, in Bergman, K.M.,
Limestone Cyclothem, Stainmore, UK: Geological Society of London, and Snedden, J.W., eds., Isolated Marine Sandbodies: Sequence Strati-
Journal, v. 151, p. 681–695. graphic Analysis and Sedimentologic Interpretation: SEPM, Special
ANTIA, E., FLEMMING, B., AND WEFER, G., 1994, Transgressive facies sequence Publication 64, p. 85–93.
of a high energy, wave-tide-storm-influenced shoreface: a case study BERNARD, H.A., LEBLANC, R.J., AND MAJOR, C.F., 1963, Recent and Pleis-
of the East Frisian barrier islands (Southern North Sea): Facies, v. 30, tocene geology of southeast Texas: Geology of Gulf Coast and central
p.15–24. Texas guidebook: Houston Geological Society, p. 175–224.
BERG, R.R., AND DAVIES, D.K., 1968, Origin of Lower Cretaceous Muddy BOYD, R., DALRYMPLE, R.W., AND ZAITLIN, B.A., 1992, Classification of clastic
Sandstone at Bell Creek Field, Montana: American Association Petro- coastal depositional environments: Sedimentary Geology, v. 80, p.
leum Geologists, Bulletin, v. 52, p. 1888–1898. 139–150.
326 H. EDWARD CLIFTON

OFFSHORE NEARSHORE
Swash
Breaker Reforming zone
zone zone

Symmetric Asymmetric Bar crest Bar trough Inner


ripples ripple (plane bed) facies planar
(bioturbated) facies (bioturbated) facies

OFFSHORE NEARSHORE

FIG. 49.—Sedimentary structural facies in a barred low-energy fine-grained nearshore under fair-weather conditions (example, Padre
Island, Texas, U.S.A., after Hill and Hunter, 1976). Sediment in the longshore trough and seaward from the bar is intensely
bioturbated.

BRUUN, P., 1962, Sea level rise as a cause of shore erosion: American Society Gilbert-type delta: Alta delta, Norway, in Colella, A., and Prior, D.B.,
of Civil Engineers, Proceedings, Journal of Waterways and Harbors, eds., Coarse-Grained Deltas: International Association of Sedimen-
v. 88, no. WW1, p. 117–130. tologists, Special Publication 10, p. 155–168.
CHIOCCI, F.C., AND CLIFTON, H.E., 1991, Gravel-filled gutter casts in CURRAY, J.R, EMMEL, F.J., AND CRAMPTON, P.J.S., 1969, Holocene history of
nearshore facies—indicators of ancient shoreline trend, in Osborne, a strandplain, lagoonal coast, Nayarit, Mexico, in Castañares, A.A.,
R.H., ed., From Shoreline to Abyss; Contributions in Marine Geol- and Phleger, F.B., eds., Coastal Lagoons, A Symposium: Mexico,
ogy in Honor of Francis Parker Shepard: SEPM, Special Publication Universi-dad Nacional Autónoma, p. 63–100.
46, p. 67–76. DALRYMPLE, R.W., AND HOOGENDOORN, E.L., 1997, Erosion and deposition
CLIFTON, H.E., 1976, Wave-formed sedimentary structures—a conceptual on migrating shoreface-attached ridges, Sable Island, Eastern Canada:
model, in Davis, R.A., and Ethington, R.L., eds., Beach and Nearshore Geoscience Canada, v. 24, p. 25–36.
Processes: Society of Economic Paleontologists and Mineralogists, DAVIDSON-ARNOTT, R.G.D., AND GREENWOOD, B., 1976, Facies relations on
Special Publication 24, p. 126–148. a barred coast, Kouchibouguac Bay, New Brunswick, Canada, in
CLIFTON, H.E., 1981, Progradational sequences in Miocene shoreline de- Davis, R.A., and Ethington, R.L., eds., Beach and Nearshore Pro-
posits, southeastern Caliente Range, California: Journal of Sedimen- cesses: Society of Economic Paleontologists and Mineralogists, Spe-
tary Petrology, v. 51, p. 165–184. cial Publication 24, p. 149–168.
CLIFTON, H.E, 2003, Coastal sedimentary facies, in Middleton, G.V., ed., DAVIS, R.A., JR., 1994, Barrier island systems—a geologic overview, in
Encyclopedia of Sediments and Sedimentary Rocks: Boston, Kluwer Davis, R.A., Jr., ed., Geology of Holocene Barrier Island Systems:
Academic Publishers, p. 149–157. New York, Springer-Verlag, p. 1–46.
CLIFTON, H.E., HUNTER, R.E., AND PHILLIPS, R.L., 1971, Depositional struc- DICKINSON, K.A., BERRYHILL, H.L., JR., AND HOLMES, C.W., 1972, Criteria for
tures and processes in the nonbarred high-energy nearshore: Journal recognizing ancient barrier coastlines, in Rigby, J.K., and Hamblin,
of Sedimentary Petrology, v. 41, p. 651–670. W.K., eds., Recognition of Ancient Sedimentary Environments:
CLIFTON, H.E., AND THOMPSON, J.K., 1978, Macaronichnus segregatis: a feed- Society of Economic Paleontologists and Mineralogists, Special
ing structure of shallow marine polychaetes: Journal of Sedimentary Publication 16, p. 192–214.
Petrology, v. 48, p. 1293–1302. Dingler, J.R., and Clifton, H.E., 1994, Barrier systems of California,
COLELLA, A., 1988a, Gilbert-type fan deltas in the Crati Basin (Pliocene– Oregon and Washington, in Davis, R.A., Jr., ed., Geology of Ho-
Holocene, southern Italy), in Colella, A., ed., Excursion Guidebook, locene Barrier Island Systems: New York, Springer-Verlag, p. 115–
International Workshop on Fan Deltas, Calabria, Italy, p. 19–77. 165.
COLELLA, A., 1988b, Pliocene–Holocene fan deltas and braid deltas in the DRAKE, D.E., AND CACCHIONE, D.A., 1985, Seasonal variation in sediment
Crati Basin, southern Italy: a consequence of varying tectonic condi- transport on the Russian River shelf, California: Continental Shelf
tions, in Nemec, W., and Steel, R.J., eds., Fan Deltas: London, Blackie Research, v. 4, p. 495–514.
& Son Ltd., p. 50–74. GALLOWAY, W.E., AND HOBDAY, D.K., 1996, Terrigenous Clastic Deposi-
CORNER, G.D., NORDAHL, E., MUNCH-ELLINGSEN, K., AND ROBERTSEN, K.R., tional Systems; Applications to Fossil Fuel and Groundwater Re-
1990, Morphology and sedimentology of an emergent fjord-head sources, Second Edition: New York, Springer, 489 p.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 327

OFFSHORE NEARSHORE
Breaker
A
Swash
Sinusoidal Buildup zone
zone
waves zone

Symmetric Asymmetric Lunate Inner


ripples ripple megaripple planar
(bioturbated) facies facies facies

OFFSHORE NEARSHORE
Swash B
Breaker Reforming zone
zone zone

Symmetric Asymmetric Bar-trough Inner


ripples ripple facies planar
(bioturbated) facies (large gravel facies
ripples)

FIG. 50.—Coastal facies on an non-barred and barred nearshores on the southeastern coast of Spain. A) Upper-shoreface profile on
a non-barred nearshore. Small (0.5 m) waves break directly on the edge of the foreshore. Lunate megaripples occur just seaward
of the beach foreshore. B) Upper-shoreface profile on a barred nearshore, southeastern coast of Spain. Large 2-D megaripples
occur in gravel at base of beach. Bar is composed of fine- to medium-grained sand.

GLAESER, J.D., 1978, Global distribution of barrier islands in terms of HARMS, J.C., MACKENZIE, D.B., AND MCCUBBIN, D.G., 1965, Depositional
tectonic setting: Journal of Geology, v. 86, p. 283–297. environment of the Fox Hills sandstones near Rock Springs, Wyoming,
GREENWOOD, B., AND SHERMAN, D.J., 1986, Hummocky cross-stratification in DeVoto, R.H., and Bitter, R.K., Sedimentation of Late Cretaceous
in the surf zone: flow parameters and bedding genesis: Sedimentol- and Tertiary Outcrops, Rock Springs, Uplift: Wyoming Geological
ogy, v. 33, p. 33–45. Association Guidebook, Nineteenth Field Conference, p. 113–130.
HADLEY, D.F., AND ELLIOTT, T., 1993, The sequence stratigraphic signifi- HARMS, J.C., SOUTHARD, J.B., SPEARING, D.R., AND WALKER, R.G., 1975, Depo-
cance of erosive-based shorefaces in the Cretaceous Mesaverde Group sitional Environments as Interpreted from Primary Sedimentary
of northwestern Colorado, in Posamentier, H.W., Summerhayes, Structures and Stratification Sequences: Society of Economic Paleon-
C.P., Haq. B.L., and Allen, G.P., eds., Sequence Stratigraphy and tologists and Mineralogists, Short Course no. 2, 161 p.
Facies Associations: International Association of Sedimentologists, HARMS, J.C., SOUTHARD, J.B., AND WALKER, R.G., 1982, Structures and Se-
Special Publication 18, p. 521–535. quences in Clastic Rocks: Society of Economic Paleontologists and
HAMPSON, G.W., AND STORMS, J.E.A., 2003, Geomorphological and se- Mineralogists, Short Course no. 9, p. 7-1–7-22.
quence stratigraphic variability in wave-dominated, shoreface–shelf HART, B.S., AND PLINT, A.G., 1995, Gravelly shoreface and beach deposits,
parasequences: Sedimentology v. 50, p. 667–701. in Plint, A.G., ed., Sedimentary Facies Analysis; A Tribute to the
328 H. EDWARD CLIFTON

A
C
older eolian dunes
2 2 fine sand,
root structures
planar parallel
1 laminated medium to B planar parallel 1 planar parallel
coarse sand laminated fine
laminated medium to
coarse sand sand
0m 0m 0m
crossbedded
-1 -1 medium to coarse -1
sand and beds or bioturbated
lenses of gravel fine sand
-2 -2 -2

-3 -3 bioturbated fine -3 bioturbated


crossbedded
sand and fine fine sand
medium to coarse
gravel and sets of
sand and beds or
-4 -4 -4
lenses of gravel hummocky
cross-
-5
stratification
-5

-6 -6
bioturbated
very fine
-7 -7 sand
bioturbated fine sand
and fine gravel
-8

older eolian dunes

FIG. 51.—Comparison of stratal successions illustrating effects of ambient wave energy and grain size. A) Pleistocene terrace
deposits, Monterey Bay, California, U.S.A., composed of medium to coarse, pebbly sand deposited in a high-energy setting
(fair-weather waves 1–2 m high; tidal range 2–3 m). B) Pleistocene terrace deposits, Mediterranean coast, southeastern Spain,
composed of coarse, pebbly sand and deposited in a low-energy setting (fair-weather waves 0.2–0.5 m high; tidal range > 1 m).
C) Eocene Jackson Group, central Texas, U.S.A., composed of fine to very fine sand and deposited in a low-energy setting (fair-
weather waves 0.2–0.5 m high; tidal range > 1 m). The Pleistocene deposit in southeastern Spain closely resembles the texturally
similar Pleistocene deposit of Monterey Bay, despite the pronounced difference in energy regime. The succession in Spain
differs markedly from that in the fine sandy deposits of the Jackson Group, which formed under a very similar energy regime.
This comparison illustrates the need to factor in the effect of grain size when interpreting paleo–wave energy. The primary
difference between the two coarse-grained deposits is that the low-energy succession is significantly thinner that that formed
under high-energy conditions, as a consequence of the deeper wave base of the high-energy California deposit.

Research and Teaching of Harold G. Reading: International Associa- Padre Island, Texas, in Davis, R.A., and Ethington, R.L., eds., Beach
tion of Sedimentologists, Special Publication 22, p. 75–99. and Nearshore Processes: Society of Economic Paleontologists and
HAYES, M.O., AND SEXTON, 1989, Modern depositional environments, Mineralogists, Special Publication 24, p. 169–187.
South Carolina: 29th International Geological Congress, 20–25 July, HOOGENDOORN, E.L., AND DALRYMPLE, R.W., 1986, Morphology, lateral
1989, Washington D.C., American Geophysical Union, Fieldtrip migration, and internal structures of shoreface-connected ridges,
Guidebook T371, 85 p. Sable Island, Nova Scotia, Canada: Geology, v. 14, p. 400–403.
HETTINGER, R.D., MCCABE, P.J., AND SHANLEY, K.W., 1994, Detailed facies HOWARD, J.D., 1966, Characteristic trace fossils in Upper Cretaceous
anatomy of transgressive and highstand systems tracts from the sandstones of the Book Cliffs and Wasatch Plateau: Utah Geological
Upper Cretaceous of southern Utah, U.S.A., in Weimer, P., and and Mineralogical Survey, Bulletin 8, p. 35–53.
Posamentier, H.W., eds., Siliciclastic Sequence Stratigraphy—Recent HOWARD, J.D., 1972, Trace fossils as criteria for recognizing shorelines in
Developments and Applications: American Association of Petroleum the stratigraphic record, in Rigby, J.K., and Hamblin, W.K., eds.,
Geologists, Memoir 58, p. 235–257. Recognition of Ancient Sedimentary Environments: Society of Eco-
HILL, G.W., AND HUNTER, R.E., 1976, Interaction of biological and geologi- nomic Paleontologists and Mineralogists, Special Publication 16, p.
cal processes in the beach and nearshore environments, northern 215–225.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 329

A A

B B

FIG. 52.—Pleistocene terrace deposits, Manresa Beach, California,


U.S.A. A) Cross-bedded sand and gravel (upper shoreface FIG. 53.—Pleistocene terrace deposits, southeastern Spain. A)
facies) exposed several meters above the base of a shallowing- Upper part of a section. Hammer head rests on contact be-
up section. B) Base of shallowing-up section. Note scattered tween cross-bedded sand and gravel (inner shoreface) and
pebbles in fine sand (lower-shoreface facies) above machete upward-fining flat-bedded pebbly sandstone (beach fore-
handle. Reddish sand with high-angle foresets at base of shore). B) Lower part of section. Cross-bedded pebbly sand in
photo is older eolian dune sand. upper part of photograph (upper shoreface) overlies bur-
rowed sand with scattered small pebbles and pebble stringers
(lower shoreface).
HOWARD, J.D., FREY, R.W., AND REINECK, H.-E., 1972, Georgia coastal region,
Sapelo Island, U.S.A.: sedimentology and biology, I. Introduction:
Senckenbergiana Maritima, v. 4, p. 3–14. KRAFT, J.C, BIGGS, R.B., AND HALSEY, S.D., 1973, Morphology and vertical
HOWARD, J.D., AND REINECK, H.-E., 1981, Depositional facies of high-energy sedimentary sequence models in Holocene transgressive barrier sys-
beach to offshore sequence: comparison with the low-energy se- tems, in Coates, D.R., ed., Coastal Geomorphology: Binghamton,
quence: American Association of Petroleum Geologists, Bulletin, v. N.Y., State University of New York, Publications in Geomorphology,
65, p. 807–830. p. 321–354.
HOYT, J.H., WEIMER, R.J., AND VERNON, J.H., 1964, Late Pleistocene and KOMAR, P.D., 1976, Beach Processes and Sedimentation: Englewood Cliffs,
Recent sedimentation, Central Georgia Coast, U.S.A., in Van Straaten, New Jersey, Prentice-Hall, Inc., 429 p.
L.M.J.U., ed., Sixth International Geologic Congress, The Nether- KOMAR, P.D., AND MILLER, M.C., 1973, The threshold of movement under
lands and Belgium, 1963, Proceedings, p. 170–176. oscillatory water waves: Journal of Sedimentary Petrology, v. 45, p.
HUNTER, R.E., CLIFTON, H.E., AND PHILLIPS, R.L., 1979, Depositional struc- 697–703.
tures and processes in oblique bar-rip channel systems, southwestern KOWALEWSKY, D.B., 1982, Wave-built terrace, in Schwartz, M.L., ed., Ency-
Oregon: Journal of Sedimentary Petrology, v. 49, p. 711–726. clopedia of Beaches and Coastal Environments: Stroudsburg, Penn-
INMAN, D.L., AND BOWEN, A.J., 1963, Flume experiments on sand transport sylvania, Hutchison Ross Publishing Company, 855 p.
by waves and currents: American Society of Civil Engineers, 8th LE ROUX, J.P., AND ELGUETA, S., 1997, Paralic parasequences associated with
Conference on Coastal Engineering, Proceedings, p. 137–150. Eocene sea-level oscillations in an active margin setting: Trihueco
JOHNSON, D.W., 1919, Shore Processes and Shoreline Development: New Formation of the Arauco Basin, Chile: Sedimentology, v. 110, p. 257–
York, John Wiley, 884 p. 276.
KERN, J.P., AND WARME, J.E., 1974, Trace fossils and bathymetry of the MACEACHERN, J.A., AND PEMBERTON, S.G., 1992, Ichnological aspects of
Upper Cretaceous Point Loma Formation, San Diego, California: Cretaceous shoreface succession and shoreface variability in the
Geological Society of America, Bulletin, v. 85, p. 893–900. Western Interior Seaway of North America, in Pemberton, S.G., ed.,
330 H. EDWARD CLIFTON

Applications of Ichnology to Petroleum Exploration: SEPM, Core American Association of Petroleum Geologists, Memoir 31, p. 247–
Workshop 17, p. 57–84. 279.
MACEACHERN, J.A., ZAITLIN, B.A., AND PEMBERTON, S.G., 1999, A sharp-based MORTON, R.A., 1994, Texas barriers, in Davis, R.A., Jr., ed., Geology of
sandstone of the Viking Formation, Joffre Field, Alberta, Canada: Holocene Barrier Island Systems: New York, Springer-Verlag, p. 75–
criteria for recognition of transgressively incised shoreface com- 114.
plexes: Journal of Sedimentary Research, v. 69, p. 876–892. NEMEC, W., 1990, Aspects of sediment movement on steep delta faces, in
MAEJIMA, W., 1993, prograding gravelly shoreline deposits in the Early Colella, A., and Prior, D.B., eds., Coarse-Grained Deltas: International
Cretaceous Yuasa Formation, western Kii Peninsula, southwest Ja- Association of Sedimentologists, Special Publication 10, p. 29–73.
pan: Geological Society of Japan, Journal, v. 89, p. 645–660. NIEDORODA, A.W., SWIFT, D.J.P., HOPKINS, T.S., AND MA, C.-M., 1984, Shoreface
MALE, W.H., 1992, The sedimentology and ichnology of the Lower Creta- morphodynamics on wave-dominated coasts: Marine Geology, v. 60,
ceous (Albian) Bluesky Formation in the Karr area of west-central p. 331–354.
Alberta, in Pemberton, S.G., ed., Applications of Ichnology to Petro- NITTROUER, C.A., DEMASTER, D.J., KUEHL, S.A., AND MCKEE, B.A., 1986,
leum Exploration: SEPM, Core Workshop 17, p. 33–55. Association of sand with mud deposits accumulating on continental
MCBRIDE, R.A., AND MOSLOW, T.F., 1991, Origin, evolution, and distribution shelves, in Knight, R.J., and McLean, J.R., eds., Shelf Sands and
of shoreface sand ridges, Atlantic inner shelf, USA: Marine Geology, Sandstones: Canadian Society of Petroleum Geologists, Memoir 11, p.
v. 97, p. 57–85. 17–25.
MIDDLETON, G.V., 1973, Johannes Walther’s Law of the Correlation of PEMBERTON, S.G., VAN WAGONER, J.C., AND WACH, G.D., 1992, Ichnofacies of
Facies: Geological Society of America, Bulletin, v. 84, p. 979–988. a wave-dominated shoreline, in Pemberton, S.G., ed., Applications of
MCCUBBIN, D.G., 1982, Barrier island and strand-plain facies, in Sholle, Ichnology to Petroleum Exploration: SEPM, Core Workshop 17, p.
P.A., and Spearing, D., eds., Sandstone Depositional Environments: 339–382.
PATTISON, S.A.J., AND WALKER, R.G., 1992, Deposition and interpretation of
long, narrow sandbodies underlain by a basinwide erosional surface:

A Cardium Formation, Cretaceous, western interior seaway, Alberta:


Journal of Sedimentary Petrology, v. 62, p. 292–309.
PLINT, A.G., 1988, Sharp-based shoreface sequences and “offshore bars” in
the Cardium Formation of Alberta: their relationship to relative
changes in sea level, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.G.,
Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea Level
Changes: An Integrated Approach: SEPM, Special Publication 42, p.
357–370.
PLINT, A.G., 1991, High-frequency relative sea level oscillations in Upper
Cretaceous shelf clastics of the Alberta foreland basin: possible evi-
dence for a glacio-eustatic control?, in MacDonald, D.I.M., ed., Sedi-
mentation, Tectonics and Eustasy: International Association of Sedi-
mentologists, Special Publication 12, p. 409–428.
POSAMENTIER, H.W., ALLEN, G.P., JAMES, D.P., AND TESSON, M., 1992, Forced
regressions in a sequence stratigraphic framework: concept, examples,
and exploration significance: American Association of Petroleum
Geologists, Bulletin, v. 76, p. 1687–1709.
B POSAMENTIER, H.W., AND CHAMBERLIN, C.J., 1993, Sequence stratigraphic
analysis of Viking Formation lowstand beach deposits at Joarcam
Field, Alberta, Canada, in Posamentier, H.W., Summerhayes, C.P.,
Haq, B.L., and Allen, G.P., eds., Sequence Stratigraphy and Facies
Associations: International Association of Sedimentologists, Special
Publication 18, p. 469–485.
POSTMA, G., 1990, Depositional architecture and facies of river and fan
deltas: a synthesis, in Colella, A., and Prior, D.B., eds., Coarse-Grained
Deltas: International Association of Sedimentologists, Special Publi-
cation 10, p. 13–28.
POSTMA, G., 1984, Slumps and their deposits on fan delta fronts: Geology,
v. 12, p. 27–30.
POSTMA, G., BABIC, L., ZUPANIC, J., AND RØE, S.-L.,1988, Delta-front failure
and associated bottom sets in a marine, gravelly Gilbert-type delta, in
Nemec, W., and Steel, R.J., eds., Fan Deltas: London, Blackie & Son
Ltd. p. 91–102.
READING, H.G., 1996, Sedimentary Environments; Processes, Facies and
FIG. 54.—A) Exposure of the Jackson Group in a uranium pit, west Stratigraphy, Third Edition: London, Blackwell Scientific Ltd., 688 p.
Texas, U.S.A. A) Upper part of section. Hammer head rests on REINECK, H.-E., 1963, Sedimentgefüge im Bereich der südlichen Nordsee:
contact between flat-bedded sand (upper foreshore) and gen- Senckenbergische Naturforschende Gesellschaft, Abhandlung, v.
tly inclined large-scale foresets with root structures 505, 138 p.
(backshore). Bioturbated sand in lower half of photo repre- REINECK, H.-E., AND SINGH, I.B, 1973, Depositional Sedimentary Environ-
sents upper shoreface. B) Lower part of section. Hammer is on ments—with Reference to Terrigenous Clastics: Berlin, Springer-
hummocky cross-stratified set in lower-shoreface or inner- Verlag, 439 p.
shelf facies. Bioturbated interval in upper part of photo is RIGBY, J.K., AND HAMBLIN, W.K., 1972, Recognition of Ancient Sedimentary
upper shoreface. Environments: Society of Economic Paleontologists and Mineralo-
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 331

Lithology Facies Paleocurrents


Unbedded siltstone INNER SHELF
Transgressive lag
Poorly sorted conglomerate,
cross-bedded sandstone FLUVIAL
A
Variegated mudstone
Structureless medium- BACKSHORE
grained sandstone
Planar-bedded FORESHORE B
medium-grained sandstone

C
Cross-bedded
coarse sandstone, UPPER D
pebble lenses and SHOREFACE
layer
E
Sharp contact
Bedded fine sandstone,
scattered small pebbles, LOWER F
burrows SHOREFACE

5m Unbedded fine sandstone


INNER SHELF

Unbedded siltstone

0 Transgressive lag
Variegated mudstone FLUVIAL

FIG. 55.—Generalized stratigraphic succession through middle Miocene shoreline deposits in the Caliente Range, California, U.S.A.
(after Clifton, 1981). Note sharp contact at base of upper-shoreface facies. Paleocurrent roses: A) Fluvial sandstone; B) planar
laminae of the foreshore; C) general summation of all (124) cross-bedding measurements within the upper shoreface facies; D)
large-scale (> 1 m) cross-bedding in the upper-shoreface facies (inferred to represent bars); E) cross-bedding in sandstone beds
immediately above the basal upper-shoreface contact (inferred to represent rip-channel deposits; F) foresets in isolated gravel
ripples in the lower-foreshore facies (inferred to represent wave ripples and the direction of wave passage).

gists, Special Publication 16, 340 p. SCHWARTZ, M.L., 1982, Beach processs, in Schwartz, M.L., ed., Encyclope-
RINE, J.M., AND GINSBURG, R.N., 1985, Depositional facies of a mud shoreface dia of Beaches and Coastal Environments: Stroudsburg, Pennsylva-
in Surinam, South America—a mud analog to sandy shallow-marine nia, Hutchison Ross Publishing Company, p. 153–157.
deposits: Journal of Sedimentary Petrology, v. 55, p. 633–652. SELLEY, R.C., 1969, Ancient Sedimentary Environments; A Brief Survey:
RINE, J.M., TILLMAN, R.W., CULVER, S.J., AND SWIFT, D.J.P., 1991, Generation Ithaca, New York, Cornell University Press, 237 p.
of late Holocene sand ridges on the middle continental shelf of New SHELTON, J.W., 1965, Trend and genesis of lowermost sandstone unit of
Jersey, USA—evidence for formation in a mid-shelf setting based on Eagle Sandstone, Billings, Montana: American Association Petro-
comparisons with a nearshore ridge, in Swift, D.J.P., Oertel, G.F., leum Geologists, Bulletin, v. 49, p. 1385–1397.
Tillman, R.W., and Thorne, J.A., eds., Shelf Sand and Sandstone SHELTON, J.W., 1967, Stratigraphic models and general criteria for recogni-
Bodies: Geometry, Facies and Sequence Stratigraphy: International tion of alluvial, barrier bar, and turbidity current sand deposits:
Association of Sedimentologists, Special Publication 14, p. 395–423. American Association of Petroleum Geologists, Bulletin, v. 51, p.
RYER, T.A., 1977, Patterns of Cretaceous shallow-marine sedimentation, 2441–2461.
Coalville and Rockport areas, Utah: Geological Society of America, SHEPARD, F.P., AND MOORE, D.G., 1955, Central Texas Coast sedimentation:
Bulletin, v. 88, p. 177–188. characteristics of sedimentary environment, Recent history, and di-
SAUNDERS, T., AND PEMBERTON, S.G., 1986, Trace fossils and sedimentology agenesis: American Association Petroleum Geologists, Bulletin, v. 39,
of the Appaloosa Sandstone: Bearpaw–Horseshoe Canyon Forma- p. 1463–1593.
tion transition, Dorothy, Alberta: Canadian Society of Petroleum SHEPARD F.P., PHLEGER, F.B., AND VAN ANDEL, T.H., 1960, Recent Sediments,
Geologists, Field Trip Guidebook, 117 p. Northwest Gulf of Mexico: Tulsa, Oklahoma, American Association
332 H. EDWARD CLIFTON

A
B

D
A
F C

E
N

100 m

B
STACKED PROGRADING BARRED SHORELINE DEPOSITS
MIOCENE, CALIENTE RANGE, CALIFORNIA

TSE

TSE
NONMARINE

UPPER SHOREFACE
TSE = Transgressive
LOWER SHOREFACE
10 m surface of
erosion SHELF
BASE OF TROUGH
0 100 m

FIG. 56.—A) Inferred coastal bar–trough systems responsible for producing the middle Miocene shoreline succession in the Caliente
Range, California, U.S.A. Shoreline trend was developed from independent evidence (after Clifton, 1981). B) Inferred stacking
pattern of the shoreline deposits shown in Part A. Individual parasquences, separated by transgressive surfaces of erosion,
contain internal erosional surfaces and sharp-based upper-shoreface deposits generated by prograding bar–trough systems.

of Petroleum Geologists, 394 p. Special Publication 64, p. 321–356.


SHIPP, R.C., 1984, Bedforms and depositional structures of a barred SWIFT, D.J.P., AND PARSONS, B.S., 1999, Shannon Sandstone of the Powder
nearshore system, eastern Long Island, New York: Marine Geology, River Basin: Orthodoxy and revisionism in stratigraphic thought, in
v. 60, p. 235–259. Bergman, K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies:
SNEDDEN, J.W., AND BERGMAN, K.M., 1999, Isolated shallow marine sand Sequence Stratigraphic Analysis and Sedimentologic Interpretation:
bodies: deposits for all interpretations, in Bergman, K.M., and Snedden, SEPM, Special Publication 64, p.85–93.
J.W., eds., Isolated Marine Sandbodies: Sequence Stratigraphic Analy- SVERDRUP, H.U., JOHNSON, M.V., AND FLEMING, R.H, 1942, The Oceans:
sis and Sedimentologic Interpretation: SEPM, Special Publication 64, Englewood Cliffs, New Jersey, Prentice-Hall, Inc., 1087 p.
p. 1–11. SWIFT, D.J.P., THORNE, J.A., AND OERTEL, G.F., 1986, Fluid processes and sea-
SNEDDEN, J.W., TILLMAN, R.W., KREISA, R.D., SCHWELLER, W.J., CULVER, S.J., floor response on a modern storm-dominated shelf: middle Atlantic
AND WINN, R.D., JR., 1994, Stratigraphy and genesis of a modern shelf of North America. Part II: Response of the shelf floor, in Knight,
shoreface-attached sand ridge, Peahala Ridge, New Jersey: Journal of R.J., and McLean, J.R., eds., Shelf Sands and Sandstones: Canadian
Sedimentary Research, v. B64, p. 560–581. Society of Petroleum Geologists, Memoir 11, p. 191–211.
SUTER, J.R., AND CLIFTON, H.E., 1999, The Shannon Sandstone and isolated VAN DE MEENE, J.W.H., BOERSMA, J.R., AND TERWINDT, J.H.J., 1996, Sedimen-
linear sand bodies: Interpretations and realizations, in Bergman, tary structures of combined flow deposits from the shoreface-con-
K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies: Sequence nected ridges along the central Dutch coast: Marine Geology, v. 131,
Stratigraphic Analysis and Sedimentologic Interpretation: SEPM, p. 151–175.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 333

1
BEACH
Fine sand,
planar lamination FORESHORE
0m

-1 UPPER
Fine to medium sand,
longshore or seaward SHOREFACE
crossbedding; -2 (BAR TROUGH
Inclined planar AND/OR
lamination
-3 RIP CHANNEL)

-4 Sharp-based ss 1

Fine to very fine sand, -5


SCS, HCS, LOWER SHOREFACE
Bioturbation TO
-6
INNER SHELF
-7
Sharp-based ss 2
Interbedded fine sand
and shale, HCS in sand, -8
bioturbation
-9

Shale, interbedded
very fine sandstone, -10
bioturbation INNER SHELF

FIG. 57.—Generalized shoreface succession on a prograding fine-grained, non-pebbly, moderate- to low-energy coast (Cretaceous
Blackhawk Fm., Book Cliffs, Utah, U.S.A.). Two sharp-based sandstones are shown: #1, produced by prograding bar–trough
systems, and #2, produced by the amalgamation of storm sandstones.

VAN STRAATEN, L.M.J.U., 1959, Minor structures of some recent littoral and Rocky Mountain area, in Peterson, J.A., and Osmond, J.C., eds.
neritic sediments: Geologie en Mijnbouw, v. 21, p. 197–216. Geometry of Sandstone Bodies: Tulsa, Oklahoma, American Associa-
WALKER, R.G., AND PLINT, A.G, 1992, Wave- and storm-dominated shallow tion of Petroleum Geologists, p. 82–97.
marine systems, in Walker, R.G., and James, N.P., eds., Facies Models: WEIMER, R.J., AND HOYT, J.H., 1964, Burrows of Callianassa major Say,
Response to Sea Level Change: Geological Association of Canada, p. geologic indicators of littoral and shallow neritic environments:
219–238. Journal of Paleontology, v. 38, p. 761–767.
WALKER, R.G., AND WISEMAN, T.R., 1995, Lowstand shorefaces, transgres- WRIGHT, L.D., CHAPPELL, J., BRADSHAW, M.P., AND COWELL, P., 1979,
sive incised shorefaces, and forced regressions: examples form the Morphodynamics of reflective and dissipative beach and nearshore
Viking Formation, Joarcam Area, Alberta: Journal of Sedimentary systems, southeastern Australia: Marine Geology, v. 32, p. 105–140.
Research, v. B65, p. 132–141.
WELLS, J.T., AND COLEMAN, J.M., 1978, Longshore transport of mud by
waves, northeastern coast of South America: Geologie en Minjbouw,
v. 57, p. 333–359.
WEIMER, R.J., 1961, Spatial dimensions of Upper Cretaceous sandstone,
334 H. EDWARD CLIFTON

TEXTURE
PEBBLY
COARSE SAND
FINE SAND
MUD

AMBIENT ENERGY SAND SUPPLY


HIGH ABUNDANT
MODERATE LIMITED
LOW

DOMINANT MORPHOLOGY
CONDITIONS
MODERATE GRADIENT
STORM VERY STEEP GRADIENT
FAIR-WEATHER
BARRED
FLOOD
BASE LEVEL
RISING
STATIC
FALLING

FIG. 58.—Parameters of open-coast settings that can influence the lithologies and stratigraphic succession of the deposit.

“STANDARD MODEL”
AMBIENT ENERGY
HIGH TEXTURE
MODERATE PEBBLY
LOW COARSE SAND
Coal FINE SAND
Planar- MUD
lamination
0
DOMINANT Cross-bedded
CONDITIONS sand
STORM
FAIR-WEATHER SCS SAND SUPPLY
FLOOD ABUNDANT
HCS LIMITED

BASE LEVEL bioturbation


RISING
STATIC MORPHOLOGY
FALLING MODERATE GRADIENT
VERY STEEP GRADIENT
10 m BARRED

FIG. 59.—”Standard” model for a prograding shoreline (Walker and Plint, 1992). Specific to a storm-dominated, moderate- to low-
energy setting with a moderate gradient in which fine sand was in somewhat limited supply and base level was static.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 335

CALIFORNIA PLEISTOCENE
Planar TEXTURE
lamination
PEBBLY
AMBIENT ENERGY 0m COARSE SAND
HIGH FINE SAND
MODERATE MUD
LOW Cross-bedded
sand and gravel

-5
DOMINANT SAND SUPPLY
CONDITIONS ABUNDANT
Scattered LIMITED
STORM small pebbles
FAIR-WEATHER
FLOOD -10
Fine sand, HCS,
wavy lamination,
bioturbation

BASE LEVEL
RISING -15 MORPHOLOGY
STATIC
FALLING MODERATE GRADIENT
VERY STEEP GRADIENT
BARRED

FIG. 60.—Prograding shoreline typical of that found in Pleistocene deposits found along the central California coast, U.S.A. Specific
to a storm-dominated, high-energy, barred setting with an abundant supply of sand, including coarse sand and gravel, under
conditions of static base level.

PLEISTOCENE SOUTHEASTERN SPAIN


TEXTURE
PEBBLY
AMBIENT ENERGY COARSE SAND
HIGH FINE SAND
MODERATE Planar MUD
LOW lamination
0m
SAND SUPPLY
ABUNDANT
DOMINANT -1 LIMITED
Cross-bedded
CONDITIONS sand and gravel
-2
STORM
FAIR-WEATHER Scattered
FLOOD small pebbles
-3

Bioturbated
-4 fine sand

BASE LEVEL MORPHOLOGY


RISING
MODERATE GRADIENT
STATIC VERY STEEP GRADIENT
FALLING
BARRED

FIG. 61.—Prograding shoreline typical of that found in Pleistocene deposits found along the southeastern coast of Spain. Specific to
a storm-dominated, low-energy, barred setting with a limited supply of sand, including coarse sand and gravel, under conditions
of static base level.
336 H. EDWARD CLIFTON

TEXAS GULF COAST


AMBIENT ENERGY TEXTURE
Root structures
HIGH PEBBLY
MODERATE Planar COARSE SAND
LOW lamination FINE SAND
MUD
0m

Bioturbated
DOMINANT fine sand
CONDITIONS SAND SUPPLY
STORM ABUNDANT
FAIR-WEATHER Alternating sets of LIMITED
FLOOD HCS and intervals of
bioturbated fine
sand
-5

BASE LEVEL Bioturbated very


fine sand
RISING MORPHOLOGY
STATIC
FALLING MODERATE GRADIENT
VERY STEEP GRADIENT
BARRED

Figure 62. Prograding shoreline typical of that found on the Texas Gulf coast. Specific to a low-energy, barred setting with an limited
supply of fine to very fine sand under conditions of static base level. Abundance of bioturbation indicates a dominance of fair-
weather conditions, although storm deposits exist in the lower part of the succession.

GILBERT DELTA
0 TEXTURE
AMBIENT ENERGY
PEBBLY
HIGH COARSE SAND
MODERATE FINE SAND
LOW MUD

Giant foresets
of sand and gravel
DOMINANT
CONDITIONS SAND SUPPLY
STORM ABUNDANT
FAIR-WEATHER LIMITED
FLOOD
10 m

BASE LEVEL MORPHOLOGY


RISING MODERATE GRADIENT
STATIC VERY STEEP GRADIENT
FALLING BARRED

FIG. 63.—Prograding shoreline in a setting where coarse sediment is introduced into a setting of low to very low energy. Although
the supply of gravel may be substantial, not enough sand enters the system to develop a offshore profile in equilibrium with the
waves. Generally associated with steep gradients and, as shown here, under conditions of static base level.
A REEXAMINATION OF CLASTIC-SHORELINE FACIES MODELS 337

FORCED REGRESSION
AMBIENT ENERGY TEXTURE
HIGH PEBBLY
MODERATE COARSE SAND
LOW Planar lamination FINE SAND
MUD
0

Cross-bedded
fine sand
SCS

DOMINANT SAND SUPPLY


CONDITIONS ABUNDANT
LIMITED
STORM
FAIR-WEATHER Mudstone interclasts
FLOOD
gutter casts

Shelf mudstone
BASE LEVEL
RISING
STATIC MORPHOLOGY
FALLING LOW GRADIENT
10 m VERY STEEP GRADIENT
BARRED

FIG. 64.—Forced-regression model for a prograding shoreline (Plint, 1988). Specific to a storm-dominated, moderate- to low-energy
setting with a low gradient in which fine sand was in somewhat limited supply and base level was falling (or had fallen).

PROGRADING BARRED SHOREFACE


Variegated mudstone TEXTURE
AMBIENT ENERGY
Structureless medium
PEBBLY
HIGH COARSE SAND
MODERATE grained sandstone
FINE SAND
LOW Planar-bedded MUD
medium grained sandstone

Cross-bedded
SAND SUPPLY
DOMINANT coarse sandstone, ABUNDANT
CONDITIONS pebble lenses and LIMITED
layer
STORM
FAIR-WEATHER
FLOOD
Erosional surface
Bedded fine sandstone,
scattered small pebbles,
burrows

BASE LEVEL Unbedded fine sandstone


RISING
STATIC MORPHOLOGY
FALLING 5m
Unbedded siltstone
MODERATE GRADIENT
VERY STEEP GRADIENT
BARRED

FIG. 65.—Prograding shoreline succession like that found in middle Miocene deposits in the Caliente Range, California, U.S.A.
Specific to a storm-dominated, moderate-energy, barred setting with an abundant supply of sand, including coarse sand and
gravel, under conditions of static sea level.
338 H. EDWARD CLIFTON
FACIES MODELS REVISITED: CLASTIC SHELVES 339

FACIES MODELS REVISITED: CLASTIC SHELVES

JOHN R. SUTER
ConocoPhillips, Sedimentary Systems, Subsurface Technology, P.O. Box 2197, Houston, Texas 77252–2197, U.S.A.
john.r.suter@conocophillips.com

ABSTRACT: A consistent, commonly accepted facies model for clastic continental shelf deposits has proven elusive. Sand ridges of various
types are ubiquitous on Quaternary shelves, but recognition of ancient examples has been intermittent, as models emerge, evolve, and
undergo relative degrees of acceptance and rejection. Continental shelves lie in the zone between highstand and lowstand shorelines, and
are greatly affected by sea-level changes. Shelf deposits per se are thus limited to and are most volumetrically significant during periods
of high or rising base level, the transgressive and highstand systems tracts. The continental shelf is roughly defined from the base of the
shoreface to the shelf margin or upper continental slope. These defining depths, and consequent shelf widths, are highly variable,
depending largely on tectonic setting. Primary shelf sediments may be derived from erosional scour of preexisting deposits, biological and/
or chemical precipitates, or supplied from adjacent shorelines by direct fluvial, deltaic, or estuarine input, or by tidal flux or storms. Fine-
grained shelf sediments, constituting potential source and seal facies, are less controversial than the origin and distribution of potential
reservoir facies. Coarse-grained shelf deposits are believed to be primarily relict, reworked from shoreline and nearshore deposits laid
down during previous regressive episodes, although notable exceptions exist. These deposits can be quite complex, taking on geometries
reflecting reshaping by wave, tidal, storm, and oceanic currents, with internal stratigraphies reflecting multi-stage depositional and
erosional histories.

INTRODUCTION (Fig. 4) is a good modern example of a broad epeiric platform


within a foreland basin.
Modern continental shelves (Fig. 1) extend from the toe of the At present it is not unfair to say that a consistent, commonly
shoreface profile (Fig. 2; also Clifton, this volume) at depths from accepted facies model for clastic shelf deposits has proven elu-
about 2 to 25 m, out to the shelf–slope break, or shelf margin. This sive. Currently, there is a general reluctance to interpret sand
latter feature occurs at a variety of depths depending on tectonic bodies within marine successions as shelfal in origin (Suter and
setting and neotectonic effects, but it is usually defined by the 200 Clifton, 1999; Galloway, 2002). The recognition of fine-grained
m bathymetric contour. Shelf widths are consequently enor- shelf deposits and the understanding of their role as hydrocarbon
mously variable, ranging from as little as 2 km to as much as 1500 source and seal facies are far less equivocal. Sand ridges of
km (Fig. 1). Most nonglaciated shelves are usually exceptionally various types are ubiquitous on Quaternary shelves (Table 1), but
flat, with seaward slopes averaging considerably less than 1°. recognition of ancient examples in the subsurface or outcrop has
Approximately 30,000,000 km2, or about 6% of the earth’s surface, been intermittent, as models emerge, evolve, and undergo rela-
can be considered continental shelf at the current glacioeustatic tive degrees of acceptance and rejection (Tables 2, 3). The dynam-
highstand. ics of continental shelves can be quite complex, with many
Continental shelves are usually classified into two major currents of different types and origins interacting (Fig. 6).
geomorphologic categories (Figs. 3, 4). Submerged continental Continental shelves lie in the zone between highstand and
margins, bounded on their updip sides by the shoreface (Clifton, lowstand shorelines, and they are greatly affected by sea-level
this volume), and in deeper waters by the shelf margin, are changes (Figs. 7–11). Fluctuating baselevels intermittently create
termed pericontinental shelves (Fig. 3). Today, pericontinental shelves by drowning continental platforms during rising phases
shelves are ubiquitous as a result of the Holocene sea-level rise. and destroy shelves by exposing the same platforms during
Drowned portions of continental interiors are termed epiconti- periods of base-level fall. The resulting deposits are subjected to
nental shelves or epeiric platforms (e.g., the Arafura Sea (Fig. 4), radically different process regimes. Shelf deposits per se are thus
the North Sea, the Baltic Sea, and the Gulf of Carpentaria). At largely limited to and are most volumetrically significant during
times in the geologic past, epicontinental seas were more com- periods of high or rising base level. These conditions occur at
mon, owing to more favorable paleogeographic configurations several scales. Eustasy has varied throughout Earth’s history
and higher eustasy. (Figs. 7, 8). High or rising base levels also feature at particular
In differing tectonic settings, clastic shelves develop into periods in an accommodation cycle, i.e., during what are known
somewhat different forms (Fig. 5; Swift and Thorne, 1991). Pas- as the late lowstand, transgressive, and early highstand systems
sive, or trailing, margins have seaward-thickening sedimentary tracts in sequence stratigraphic terminology (Fig. 9). Shelf depos-
wedges, typically forming broad pericontinental shelves, of which its reflect this complexity. Many deposits on modern shelves are
the Gulf of Mexico (Fig. 3) is an excellent example. Convergent of composite origin, beginning as shoreline facies that undergo
margins form around subduction zones, and are usually hinged varying degrees of reworking upon transgression. Morphology,
on their seaward side. Shelf areas are typically narrow, wave-cut lithofacies, and vertical successions of these deposits can be quite
platforms, but they may also be aggradational accretionary prisms. similar to those of adjacent shoreface, deltaic, or estuarine depos-
Foreland basins are formed in front of fold-and-thrust belts, with its. In addition, even the most thoroughly studied and best-
maximum rates of subsidence and deposition near the orogenic known modern shelf sands occur in a transitional environment
belt. A peripheral bulge may complicate this simple profile, that exists only at certain periods in an accommodation cycle (Fig.
which tends to produce a basinward-thinning sedimentary wedge. 9). Preservation in their current form is doubtful. As such, inter-
The Arafura Sea offshore of Papua New Guinea and Irian Jaya pretations of ancient shelf sediments based on Quaternary ex-

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 339–397.
340 JOHN R. SUTER

0 5,000 10,000 km

FIG. 1.—Satellite topographic–bathymetric image of the world (image courtesy of K. Soofi, ConocoPhillips, 2004). Continental shelves
are those portions of the ocean basins shown in light blue, comprising depths between 10 and 200 m. This designation represents
about 6% of the total area of the globe. Slopes on the continental shelves range from about 0.001° to ~ 1°, and widths vary from
less than 10 km to greater than 200 km.

Foreshore
Shoreface

Shelf

Foreshore

Shoreface

Shelf

FIG. 2.—Digital bathymetry and profile of a modern shoreface–shelf transition, from offshore of Fraser Island in northeastern
Australia (image courtesy of R. Boyd, 2004). The continental shelf begins at the change in slope at the toe of the lower shoreface.
The depth at which this occurs is highly variable and ranges from about 2 m to as much as 25 m in different parts of the world
(Clifton, this volume).
FACIES MODELS REVISITED: CLASTIC SHELVES 341

Mississippi
Valley

Houston Mississippi Delta

in e
orel
Sh Continental
Continental Shelf
Shelf
Mississippi
Mississippi
Maarg
SShheelflf M in 220000 m
rgin m Canyon
Canyon

Continental
Continental Slope
Slope

Abyssal
Abyssal Plain
Plain

00 200
200 km
km

FIG. 3.—The Gulf of Mexico is an excellent example of a submerged continental margin, or pericontinental clastic shelf. Shelves of this
kind are the most common today. The continental shelf extends from just seaward of the shoreline to the shelf margin, which lies
just landward of the 200 m contour indicated. The rugose topography of the continental slope in this area is the result of extensive
salt tectonics. (Satellite topography–bathymetry image courtesy of K. Soofi, ConocoPhillips).

amples are often suspect, and many frequently cited examples Sea as comprising clastic shelf deposits. Some major Cretaceous
have been reinterpreted to represent other depositional environ- sandstone reservoirs in the Western Interior of North America are
ments. Although it is a daunting task, this review utilizes what is thought to be shelf sand deposits. The Tocito Sandstone (e.g.,
known about Quaternary clastic shelf deposits in an attempt to Nummedal and Riley, 1999) has a cumulative oil production of
provide some consistent working parameters, if not a consistent over 160 MMBO. Hartzog Draw (e.g., Tillman and Martinsen,
facies model, for the interpretation of clastic shelf deposits in the 1987), which is dominantly a Shannon Sandstone field, has an
ancient record. estimated 400 MMBO in place. The Shannon and similar “isolated
Given this degree of complexity, the importance of clastic marine sands” (Snedden and Bergman, 1999) are the archetypal
shelf sands as hydrocarbon reservoirs is similarly uncertain (see “offshore bars.” Sadly, this particular interpretation is extremely
papers in Bergman and Snedden, 1999; Galloway 2002). Shelf controversial (see the numerous papers in Bergman and Snedden,
sands have been interpreted as the principal reservoir facies in 1999) with at least four competing alternative interpretations
numerous oil and gas fields (Table 3), although with the advent (Suter and Clifton, 1999).
of sequence stratigraphy in the 1980s many previous “offshore
bars,” or what would today be called shelf sand ridges, were SEDIMENT SUPPLY TO CLASTIC SHELVES
reinterpreted as incised valleys or forced regressive shoreface
deposits (Fig. 11; Walker and Plint, 1992; Sullivan et al, 1997; Probably the most vexing question relating to continental-
Snedden and Bergman, 1999). Recently, Posamentier (2002) inter- shelf sands and gravels concerns the sediments themselves:
preted Miocene reservoirs from Indonesia as shelf sands, and How did they get out there? Original models of continental-
Galloway (2002) interpreted Miocene deposits from the North shelf sediments used the graded-shelf concept, in which it was
342 JOHN R. SUTER

~ 10 – 200 m Water Depth

Irian Jaya
M
Moou
unnt
taaiin
n BB
eellt
t

Papua -
New Guinea
Co
as
tal
Pla
Arafura i n Fly
Delta
Sea

in
gin
Maarrg
lf M
Cape

Shheelf
York

S
Australia
Coral Sea
Gulf
Gulf of
of Carpentaria
Carpentaria

0 500 km
Australia
FIG. 4.—The Arafura Sea between Australia and Papua New Guinea–Irian Jaya is a good modern example of an epicontinental sea
(satellite topography–bathymetry image courtesy of K. Soofi, ConocoPhillips). A shelf margin does exist eastward of the Fly Delta
and the York Peninsula, but westward a broad, submerged continental platform extends some 1000 km to the Indonesian island
arc.

Passive margin
(A) Passive margin
B C

Convergent
(B) margin
Convergent margin Foreland
(C) basin
Foreland basin
Shelf Width
FIG. 5.—Variations in the tectonic setting of continental shelves (redrawn from Swift and Thorne, 1991). A) Passive margins, or trailing
margins, are pericontinental shelves (Fig. 3), with increasing rates of subsidence seaward, and consequent seaward-thickening
sedimentary deposits. B) Convergent margins form around subduction zones. Subsidence rates are greatest in the area of
subduction. Shelf areas are typically narrow, wave-cut platforms, but they may also be aggradational accretionary prisms. C)
Foreland basins also occur at convergent margins, with maximum rates of subsidence and deposition adjacent to the thrust belt.
Extensive shelf areas commonly develop.
FACIES MODELS REVISITED: CLASTIC SHELVES 343

TABLE 1.—Some Quaternary shelf sands: a non-exhaustive list of sands present on present-day continental shelves.
Many of these deposits have been written about multiple times. No effort has been made to list every pertinent
reference, but hopefully those provided will be a starting point for further examination.

Location Water Depth, m Interpretation References


NE Brazil margin 13–20 Ocean current-dominated Testa and Bosence (1999)
Yang and Sun (1988)
Yang (1989)
East China Sea 45–115 Tide-dominated
Liu et al. (1998)
Berné et al. (2002)
Gulf of Lyons > 100 Wave-dominated Berné et al. (1998)
Keller and Richards (1967)
Straits of Malacca 50–80 Tide-dominated
Emmel and Curray (1982)
Eastern Australia 20–70 Wave-dominated Roy et al. (1994)
Flemming, (1980); (1981)
Southeast Africa 35–70 Ocean current-dominated
Ramsay et al. (1994)
Atlantic Margin, North America
East Coast, USA < 20* Storm-dominated McBride and Moslow (1991)
Maryland, USA 16–30 Storm-dominated Swift and Field (1981)
Stubblefield et al. (1984)
New Jersey, USA 4–45 Storm-dominated Rine et al. (1991)
Snedden et al. (1999)
Swift et al. (1986)
New Jersey, USA 20–120 Storms and ocean currents
Goff et al. (1999, 2005)
Twichell (1983)
Georges Bank 25–50 Tide- and storm-dominated
Dalrymple et al. (1992)
Hoogendorn and Dalrymple (1986)
Sable Island Bank Storm-dominated
Dalrymple and Hoogendoorn (1997)
North Sea
Erosional lowstand
Berné et al. (1998)
Celtic Sea ≥ 100 remnants
Tide-dominated moribund Reynaud et al. (1999)
Belgian coast 10–20 Tide-dominated Berné et al. (1994)
Houbolt (1968)
Belderson et al. (1982)
Belderson (1986)
North Sea 15–40 Tide-dominated
Stride et al. (1982)
Huthnance (1982)
Davis and Balson (1992)
Gulf of Mexico, USA
Curray (1960, 1964)
Nelson and Bray (1970)
Frazier (1974)
East Texas/ Southwest Louisiana 12–29 Storm-dominated
Suter (1987)
Thomas and Anderson (1989)
Rodriquez et al. (1999)
Penland et al. (1986, 1988, 1989)
Mississippi Delta 10–50 Storm-dominated
Pope et al (1990)
Northeast 20–35 Transgressive sand sheet McBride et al. (1999)
Southwest Florida <7 Tide-dominated Davis et al. (1993)

believed that sediments should change from coarse to fine as eustatic fluctuation and had been left behind on the continental
one progressively moved from the shoreline into deeper water. shelf in the ensuing Holocene transgression. Emery (1968) coined
This reflected decreasing current and wave energy in an off- the term relict for these sediments, meaning that they were
shore direction, and the general belief that no mechanism ex- erosional remnants of previous processes and were not in equi-
isted to transport coarse sediments more than a few kilometers librium with the process environment of the modern shelf.
from the shoreline. As oceanographic studies increased in num- Further research continued to demonstrate that many shelves
ber and importance in the latter half of the 20th century, it are highly energetic, with currents capable of considerable
became apparent that neither the type nor the distribution of erosion and sediment transport. Swift et al. (1971) termed shelf
sediments on continental shelves varies smoothly from the sediments palimpsest if they were brought to a particular loca-
shoreline to the shelf margin. Coarse sands and gravels were tion by some other mechanism but achieved hydraulic equilib-
found at various depths on shelves across the globe. Curray rium with subsequent shelf processes. The realization grew that
(1960, 1964; see Figure 10), among others, realized that much of there were a number of different types of sands on continental
this sediment distribution was the result of the recent glacio- shelves, emplaced by a variety of mechanisms. Debate contin-
344 JOHN R. SUTER

TABLE 2.—Some interpreted clastic shelf sands in outcrop. Many of these deposits, in particular the Shannon Sandstone,
have been written about multiple times, and almost all, if not every one of these examples, has one or more
competing interpretations. References in this table mostly advocate the shelf interpretation;
those which represent competing viewpoints are marked by an asterisk*.

Formation Age Location Type Reference


Kazusa Group Pleistocene Boso Peninsula, Japan Current-dominated Ito and Horikawa (2000)
Spearing (1976)
Brenner (1978)
Shannon Cretaceous Wyoming, USA Storm-dominated Tillman and Martinsen (1987)
Gaynor and Swift (1988)
Bergman (1994)*
Bergman and Snedden (1999)
Jennette and Jones (1994)*
Tocito Cretaceous New Mexico, USA Tide-dominated
Nummedal and Riley (1999)
Krause et al. (1987)
Cardium Cretaceous Western Canada Storm-dominated
Plint (1988)*
Frontier Cretaceous Wyoming, USA Storm-dominated Winn (1991)
Cole and Young (1991)
Mancos B Cretaceous Utah and Colorado, USA Shelf Plume Cole et al. (1997)
Hampson et al. (1999)*
Scatter Cretaceous Northwest Canada Storm-dominated Leckie and Potocki (1998)
Niel Klinter Jurassic East Greenland Storm-dominated Dam and Surlyk (1998)
Raukely Cretaceous East Greenland Tide-dominated Surlyk and Noe-Nygaard (1991)
Duffy Mountain Cretaceous Colorado, USA Shelf plume, storm sands Boyles and Scott (1982)
Vlierzele Eocene Belgium Tide-dominated Houthuys and Gullentops (1988)
Rogn Jurassic Northern North Sea Tide-dominated Goetsen and Nelson (1992)
Armorica Quartzite Ordovician Southern Spain Storm-dominated Rey and Hidalgo (2004)
Various Silurian Tennessee, USA Storm-dominated Driese et al. (1991)
Gog Group Cambrian Western Canada Storm-dominated Hein et al. (1987)
Sandfjord Precambrian Finnmark, Norway Tide-dominated Levell (1980)
Precambrian–
Various Mid-continent Tide-dominated Various
Paleozoic

ues concerning how far and by what mechanisms sands can be Allochthonous Sediments
transported from the shoreline onto the shelf (Dalrymple and
Cummings, 2004). Sediments flowing onto the continental shelf from a river or
Sediments supplied to continental shelves and ultimately an estuary (Fig. 14) are immediately subject to a number of shelf
constituting shelf deposits fall into several broad categories. dispersal processes (Figs. 6, 15, 16). Despite decades of study, and
Allochthonous (Swift and Thorne, 1991) sands, silts, clays, and considerable debate, the mechanisms and magnitude of cross-
organic matter are those particles transported to the shelf by shelf dispersal and deposition are still not fully understood
fluvial/estuarine processes, bypassing river mouths as coastal (Nittrouer and Wright, 1994; Leeder, 1999; Dalrymple and
jets, hyperpycnal flows, hypopycnal plumes, or tidal flux. Sedi- Cummings, 2004). When sediments from a fluvial source enter a
ments are also transported from the shoreline by longshore marine receiving basin, they have a number of possibilities,
currents and by downwelling storm flows. If sediment is sup- summarized in Figures 15 and 16. Sediments entering a marine
plied at a rate greater than the creation of accommodation on the basin have three basic fates: (1) deposition nearshore; (2) trans-
shelf, a progradational or supply-dominated shelf setting results portation seaward or shore parallel as hypopycnal (buoyant)
(Fig. 12; Swift and Thorne, 1991). Autochthonous sediments are plumes; or (3) transportation seaward as hyperpycnal gravity
those developed in situ, by erosion and resuspension by waves or flows.
upwelling, by the formation of authigenic minerals such as If a fluvial system provides enough sediment to form a delta,
glauconite, or by direct creation from near-surface organic pro- bed load carried by traction currents is usually deposited close to
ductivity. Supply-dominated shelves are somewhat limited shore in the form of distributary-mouth bars (e.g., the Mississippi
in the present day, but clearly they were more common in the Delta, Fig. 14; also see Bhattacharya, this volume). Wave-domi-
ancient record. Many supply-dominated shelves are on active nated estuaries may have ebb tidal deltas that transport sediment
margins, where small mountainous rivers deliver the bulk of onto the inner shelf, while tide-dominated estuaries commonly
sediment to the world’s oceans. Accommodation-dominated, or have a series of linear bars in the transition area from estuary to
transgressive, shelves—more typical of present-day conditions inner shelf (see Boyd et al., this volume). These mechanisms
on most pericontinental margins—result when sediment supply emplacement coarse sediments onto the continental shelf, but
is insufficient to fill available space (Fig. 13; Swift and Thorne, these sands belong to their respective depositional environments
1991). and are not considered shelf deposits per se.
FACIES MODELS REVISITED: CLASTIC SHELVES 345

TABLE 3.—A non-exhaustive list of interpreted shelf sands in the subsurface. Many of these examples have been
interpreted as some other depositional environment, and several have more than one competing interpretation.
The references provided here support the shelf sand for the respective reservoir, and are meant as representative
starting points, not as a complete bibliography. We can conclude that (1) interpreting shelf sandstones in the
subsurface is not straightforward, and (2) it is problematic to use interpreted examples of shelf sands from
outcrop or the subsurface to construct a facies model for clastic shelf deposits.

Field Example Formation Age Location Reference


Mariner Anahuac-Fleming Miocene Mississippi, USA Handford and Baria (2003)
Seria Miri Miocene Sabah Atkinson et al. (1986)
Various Upper Cibulakan Miocene Indonesia, Java Sea Posamentier (2002)
None – possible CO2 Utsira Miocene North Sea Galloway (2002)
sequestration candidate

Pembina Cardium Cretaceous Alberta, Canada Krause and Nelson (1991)


Kinsale Head, Ballycotton Greensand-Gault Cretaceous UK North Sea Winn (1994)
“A” Sand
Kuparak Kuparak River Cretaceous Alaska Gaynor and Scheihing (1988)
Milk River Milk River Cretaceous Alberta/Saskatchewan, Canada Meijer-Drees and Myhr (1981)
Tom Walsh-Owen Olmos Cretaceous South Texas, USA Snedden and Jumper (1990)
Hartzog Draw Shannon Cretaceous Wyoming, USA Tillman and Martinsen (1987)
House Draw Sussex Cretaceous Wyoming, USA Berg (1975)
Higley et al. (1997)
Bisti Tocito Cretaceous New Mexico, USA Nummedal and Riley (1999)
Various Bakken Shale Devonian– Western Canada Toews et al. (2004)
Mississippian
Heidrun Fangst Group Jurassic Norwegian North Sea Harris (1989)
Whitley (1992)
Alpine Kingak Shale Jurassic Alaska Morris et al. (2000)
Burns et al. (2002)
Draugen Rogn Jurassic Norwegian North Sea Goetsen and Nelson (1992)
Troll Sognefjord, Jurassic Norwegian North Sea Gray (1987)
Fensfjord

Buoyant or hypopycnal plumes (Figs. 14–18) are less dense suspended mud and silt kept in suspension by the turbulent
than the ambient seawater. Such plumes may reach well out effluent, as well as additional turbulence from waves and tidal
onto the shelf or across the shelf break, depending on their own forcing. Concentrations reach those required for “fluid muds”
characteristics, the slope of the shelf, and the prevailing shelf (> 10 g/l; Faas, 1991) in many areas of the plume (Kineke et al.,
winds, waves, and currents. Suspended load in a hypopycnal 1996), although concentrations are generally lower throughout
plume is deposited as the plume loses momentum into the much of the plume. Fluid muds can be as much as 7 meters thick,
receiving basin. Individual grains settle out as turbulence de- and they appear to form by enhanced settling and convergence
creases, abetted by the process of flocculation, the agglomera- of near-bottom flows (Kineke et al., 1996), similar to the pro-
tion or clumping of suspended particles into grains which can cesses that occur in an estuarine turbidity maximum (see Boyd
settle more quickly. Flocculation occurs as positively charged et al., this volume). Sediments carried by the Amazon jet appear
ions in seawater interact with negatively charged clay particles. to be cyclically deposited and resuspended near the river mouth
Typically this happens when salinities exceed 1–5% (Dyer, until being carried offshore and along the shelf (Kineke et al.,
1995), which may either take place as a salt-water wedge in 1996).
rivers or estuaries or extend offshore for many kilometers Most hypopycnal plumes are turned by the Coriolis force or
(Bentley, 2003). intruding oceanic currents before reaching past midshelf depths
The plume of the Amazon River can extend offshore over 100 (Nittrouer and Wright, 1994). The rotation of the earth imparts the
km (Nittrouer et al., 1986) before turning to the northwest to apparent Coriolis force to horizontally moving objects, creating a
form a huge “river” of mud moving along the northeastern coast deflection that is oriented at right angles to the direction of
of South America (Figs. 17, 18). This transport is so significant movement. In the northern hemisphere the deflection is to the
that, despite enormous deposition of muds along the coasts of right, and in the southern hemisphere, to the left. This deflection
French Guiana, Surinam, and Guyana (Fig. 18, Wells and affects offshore-flowing marine currents, causing them to veer
Coleman, 1981; Rine and Ginsburg, 1985) substantial portions of obliquely offshore. Ultimately, should the pattern persist for
the Orinoco delta comprise mud originally sourced from the sufficient time to achieve equilibrium, the offshore flow and the
Amazon River (Aslan et al., 2003). The plume has tremendous Coriolis deflection balance, creating an essentially shore-parallel
momentum as it leaves the Amazon River mouth, laden with geostrophic, or “balance of forces”, current.
346 JOHN R. SUTER

Oceanic
Oceanic Tides
Tides Meteorological
Meteorological Density
Density

Reversing
Reversing Rotary
Rotary Longshore
Longshore Direct
Direct wind
wind

Cyclic
Cyclic Residual
Residual Wind
Wind Wind
drift
drift setup

Landward
Landward Homopycnal
Homopycnal Internal
Internal Hyperpycnal
Hyperpycnal
bottom
bottom Plumes
Plumes Waves
Waves flows
flows
FIG. 6.—The dynamics of continental shelves can be quite complicated, owing to the interaction of the various components of the
current field shown above. Ordinarily a given shelf is dominated by one process or another. All of the currents illustrated above
combine with the Coriolis force caused by the earth’s rotation to form geostrophic “balance of forces” currents (redrawn from
Leeder, 1999, and Swift, 1972).

Interglacial Periods - Higher Sea Level - increased shelf area


1 - 2.0
9 7
11 5
13

0.0 δ18 O
3 (‰)

4
10 8
12 6
Glacial Periods - Lower Sea Level - decreased shelf area 2 + 2.0

500 400 300 200 100 Present


Time (Thousands of years before present)

FIG. 7.—Oxygen-isotope stages for the past 500,000 yr (redrawn from Imbrie et al., 1984). Even-numbered stages depict glacial
periods, in which ocean water is sequestered into continental ice, resulting in lowered global sea levels and decreased areas
of continental shelves. Odd-numbered stages represent interglacial times, when melting of icecaps returns water to the world
oceans, creating higher sea levels and expanded continental shelf areas. Most previous highstands appear to have been within
a meter or two of present-day sea levels. Estimates for maximum lowstand in the last glacial period (Stage 2), range from -90
to about -120 m.
FACIES MODELS REVISITED: CLASTIC SHELVES 347

SEQUENCE SEQUENCE OXYGEN


CHRONOSTRATIGRAPHY BOUNDARIES ISOTOPES
TIME T-R FACIES
MAJOR δ18O, ‰
SEQUENCES T-R
(Ma) CYCLES 0 1 2 3 4
CYCLES
HOLOCENE 0
R R
Q PLEISTOCENE
.80
1.40
Cala 1
Io 1
Cala 2
0.8=lo 1

1.56 0.6=Cala 1
2.09
Ge 2 PGI-2
PLIOCENE

2.55
2.76
Ge 1 2.4=Ge 1 PGI-1
Pia 2
U

3.21
PPI-2
Pia 1 3.0=Pla 1 PPI-1
4.04
Za 2 3.8=Za 2
4.37
Za 1 4.2=Za 1 PZI-3
PZI-2
L

5 PZI-1
T
5.73
Me 2 5.5=Me 2 MMI-2
R
6.98
MMI-1
Tor3/Me 1 6.3=Tor 3/Me 1
UPPER

MTI-4

MTI-3
9.26
Tor 2 8.2=Tor 2 MTI-2
10
TERTIARY

MTI-1
T
MIOCENE

11.70
Ser 4/Tor 1 10.5= Ser 4/Tor1 MSI-4
R
MIDDLE

12.70
Ser 3 MSI-3
12.5= Ser 3
13.60
Ser 2 MSI-2
13.8= Ser 2
14.80 MSI-1
15 Lan 2/Ser 1
15.5= Lan 2/Ser 1

16.40
Bur 5/Lan 1 16.5= Bur 5/Lan 1 MLI-1

17.30
MBI-3
Bur 4 17.5= Bur 4

18.70
Bur 3 MBI-2
LOWER

19.50
Bur 2 MBI-1
20
20.52
Aq 3/Bur 1 21=Aq 3/Bur 1
MAI-3
22=Aq 2
22.20
Aq 2 MAI-2

23.80
T MAI-1
Ch 4/Aq 1
R 25.5=Ch 4/Aq 1
25

FIG. 8.—A portion of the eustatic cycle chart documenting periods of coastal onlap, indicating rise and fall of sea level (redrawn from
Hardenbol et al., 2000). Periods of high and/or rising base level create expanded continental shelf areas and more favorable
conditions for shelf sand deposition.

Many other present-day fluvial systems support large mud- Hyperpycnal flows (Figs. 16, 20) are denser than seawater,
laden plumes, including the Mississippi–Atchafalaya River com- and move offshore (downslope) until drag forces are sufficient
plex of North America (Figs. 14, 19, e.g., Bentley, 2003; Neill and to exceed the downslope momentum created by gravity and
Allison, 2005), the Yellow River of China (e.g., Wright et al., 1988; immersed weight. Theoretically, hyperpycnal plumes can form
Liu et al., 2004), and the Po River of Italy (Cattaneo et al., 2003). at the mouths of rivers that are able to generate sufficiently
Although such plumes are commonly associated with major deltas concentrated flows to overcome the ambient ocean density
(Fig. 14), they can produce deposits which differ from the prodelta (Mulder and Syvitski, 1995). This requires very high sediment
of classical deltaic facies models (Bentley, 2003; Cattaneo et al, concentrations, and until fairly recently it was generally ac-
2003). In some, and perhaps most, cases, sediment supply from cepted that hyperpycnal plumes issuing directly from river
numerous smaller rivers may combine to form a “subaqueous mouths were relatively rare phenomena on modern continental
delta” which cannot be traced directly to a single fluvial source shelves, because relatively steep gradients (> 0.7°) are required
(Cattaneo et al., 2003). Numerous thick successions of shelf mud in to generate and maintain them (Bentley, 2003). However, a
the ancient record may well be this type of deposit. Because plumes number of authors (e.g., Wright et al., 1988; Mulder and Syvitski,
are so commonly turned alongshore, these deposits may have 1995; Liu et al., 2004) have documented such flows, particularly
progradation directions subparallel to shelf contours, confusing from the Huang He (Yellow River) of China. Our collective
correlations and paleogeographic reconstructions. understanding of these processes remains incomplete. Parsons
348 JOHN R. SUTER

Maximum
Highstand Highstand eustatic
Rising RSL Fallling RSL rise

SL
gR
Risin

llin
Fa
Maxim

Ri
RSL

si
ng
um Flo easing she

RS
RSL
- incr

L
"R - inflection Point" -
Maximum rate of eustatic sea-level rise

ing
oding

Fall
Surfa r e a

Ri
si
ng
RS
lf a
ce Fo

n
L io
Maximum at
rm a
RSL

Melt-water Fo a r e
rma

discharge y
r lf
da s h e
ing
tion

n
u g
Fall

Bo s i n
e a
e nc c r e
u e
S eq - D
Maximum
Eustatic Fall

Falling base level “Highstand” Rising baselevel “Lowstand”


Eustatic Fall “Normal” conditions Eustatic Transgression
Forced Regression Maximum Shelf Area Retrogradation Minimum Shelf Area;
Fluvial Extension and Increasing Shelf area Subaerial exposure
Valley Incision Delta Plain aggradation Shelf Margin and
Subaerial exposure Fluvial/Estuarine Valley filling Deepwater deposition
Decreasing Shelf Area Transgressive Erosion
Shelf Sand development

FIG. 9.—Hypothetical response of to a Quaternary glacioeustatic fluctuation in the northern Gulf of Mexico (redrawn from Suter,
2003). No exact timing of different depositional styles, fluvial–estuarine responses, or significant surface formation is intended.
Such cycles are asymmetric and contain several smaller eustatic fluctuations; two are shown here, compare with Figure 7. This
has significant impact on the stratigraphic architecture of shelf and shelf-margin deposits, particularly when autocyclic processes
and climatically driven changes in discharge and sediment load are convolved with base-level forcing. A greater amount of time
is actually spent during overall falling base level and lowstand phases, during which the incised/extended fluvial drainage
network is formed. Continental shelf area expands and contracts as sea level rises and falls. Consequently, continental-shelf
deposits, and in particular shelf sands, are most significant in the late lowstand, transgressive, and early highstand systems tracts.
Note that, in a late Quaternary glacioeustatic cycle, time involved in the formation of the transgressive maximum-flooding surface
is less than that required to form the regressive sequence boundary.

et al. (2001) demonstrated that hyperpycnal flows could be ment gravity flows at much lower gradients than needed for
generated at substantially lower concentrations than previ- autosuspended flows (Fig. 20).
ously thought, through convective processes which coupled The importance of hyperpycnal flows has likely been under-
hypopycnal and hyperpycnal plumes. Bentley (2003) suggested estimated to date (Leeder, 1999), and increasing numbers of
that flocculation, coupled with acceleration of a hypopycnal authors are invoking this mechanism for deposition on the shelf
plume, could result in dense near-bed suspensions, helping to as well as in deeper water (e.g., Plink-Björklund et al., 2001;
achieve conditions for the generation of hyperpycnal flows. Johnson et al., 2001; Bentley and Nittrouer, 2003; Plink-Björklund
Several authors (e.g., Kineke et al., 1996; Kineke et al., 2000; and Steel, 2004; Edwards et al., 2005; Pattison, 2005). Sediment
Myrow and Southard, 1996; Traykovski et al., 2000; Wright et gravity flows have appeal as a mechanism for the transport of
al., 2001; Myrow et al., 2002; Bentley, 2003) suggest that turbu- sands onto the continental shelf, but there are some intriguing
lence generated by waves and currents can help to ignite sedi- caveats. Khan et al. (2005) have shown that alongshore currents
FACIES MODELS REVISITED: CLASTIC SHELVES 349

M
Louisiana iss
iss
ipp
iR
ive
r

Trinity
Shoal
LaFourche
Delta Mississippi
Maringouin Delta
Delta
St. Bernard
Ship Shoal "Outer" Delta
Southwest Louisiana Mississippi
Shoal
Shelf Sands Canyon

. 75 - 100% Sand
600 ft

0 50 Mi. 50 - 75% Sand


GULF OF MEXICO Silty Clay
0 100 Km.
CPH-060046.13

FIG. 10.—Distribution of surficial sediments on the Louisiana continental shelf of the northern Gulf of Mexico (redrawn from Frazier,
1974). Ship, Trinity, and the “Outer” Shoal are shelf sand bodies related to the transgression of abandoned Maringouin and
LaFourche complexes of the Holocene Mississippi delta (Penland et al., 1986; Penland et al., 1988). East of the modern Mississippi
Delta is another area of shelf sands, related to the transgression of the abandoned St. Bernard complex of the Mississippi Delta
(Penland et al., 1988; Penland et al., 1989). The Southwest Louisiana shelf sands, and their extension into the waters off of East
Texas, resulted from transgression of Pleistocene shoreline and coastal-plain deposits during the Holocene sea-level rise (Curray,
1960; Frazier, 1974; Suter, 1987).

FIG. 11.—Generation of shallow marine sandbodies by fluctuations in sea level (redrawn from Walker and Plint, 1992). A) Wave-
dominated shorelines form as part of a highstand systems tract (sea level I). Subsequent fall in base level causes “forced
regression,” forming a series of shoreface deposits at successively seaward positions (vertical profiles 2–3), creating “falling stage”
and lowstand (profiles 4–6) systems tracts (Sea levels II–VII). At lowstand, a prograding shoreface is reestablished (sea level VII).
B) Rising base level submerges the exposed coastal plain, expanding the continental shelf. Lowstand and falling-stage deposits
are submerged and reworked by shoreface processes. The “ravinement surface” caused by shoreface erosion truncates
underlying deposits (vertical profiles 7–12). In places, the transgressive erosion surface can be recognized only by an erosional
surface overlain by a thin (centimeter scale) transgressive lag (e.g., thin pebble layer; vertical profile 9). Although not illustrated
in this diagram, marine processes, including waves, tides, and currents, continue to rework the shoreline deposits into shelf sand
bodies, as will be developed further in this chapter. Given sufficient sediment supply, shelf muds bury the earlier sand deposits.
350 JOHN R. SUTER

Sediment Supply > Rate of Accommodation Creation: Progradational

Shelf Floor
Graded, equilibrium surface(s)
Basinward fining Slope-basin Floor
SUPPLY-DOMINATED SHELF River mouth bypassing Gravity transport
Sediment accretion
Ocean currents

FIG. 12.—A) Swift and Thorne (1991) classified shelves on the basis of the ratio of the rate of sediment supply to the rate of creation
of accommodation. Supply-dominated shelves are those in which incoming sediment exceeds the space available to be filled,
resulting in progradation. Rivers supply autochthonous sediment directly to the shelf, and deltas and regressive shoreface deposits
are common (redrawn from Swift and Thorne, 1991). B) An idealized profile and summary of processes and features of a supply-
dominated shelf setting (redrawn from Johnson and Baldwin, 1996).

can significantly alter the shape and direction of hyperpycnal sediments. Offshore flows are deflected by the Coriolis force to
flows issuing from river mouths, shifting the flow to a shore- create geostrophic, essentially shore-parallel currents, which limit
parallel configuration and limiting the amount of cross-shelf the distance to which sediments can be transported offshore
transport. Field observations and numerical models indicate that (Dalrymple and Cummings, 2004).
hyperpycnal flows are more likely in smaller, mountainous rivers Aigner and Reineck (1982) and Aigner (1985) provided a
found on convergent margins than in the larger fluvial systems detailed description of the proximal–distal trends produced by
that often characterize pericontinental settings (Syvitski et al., storm sediment transport in the German Bight (Fig. 24). Near-
2003). This leads to a conundrum in that a mechanism to transport shore, deposits are amalgamated, thicker, and more numerous,
sand onto the shelf is likelier to function in supply-dominated reflecting the fact that many more storms can impact the near-
settings, whereas most of our models and interpretations for shelf shore area. Such deposits interfinger with shoreface sediments
sands come from accommodation-dominated examples. (see Clifton, this volume) in a shoreward direction. Farther off-
Sediments are also delivered from the nearshore into shelf shore, individual beds are preserved, with decreasing lamination
environments by storms. As a storm impacts a coastal/shelf and increasing amounts of bioturbation. The distance of offshore
system (Fig. 21), water levels at the coastline are “set up,” or rise, transport is dependent upon a number of factors, including the
owing to several factors, including wind-driven frictional cou- slope of the system, the degree of setup of the water level at the
pling and water setup, wave shoaling and runup, astronomical shoreline, and the intensity and duration of the storm.
tides, and the so-called inverse barometer effect—expansion of Sediment resuspension by storm waves is an important pro-
the water at lowered atmospheric pressure. The increased water cess (Nittrouer and Wright, 1994). Once the sediment has been
level creates a hydraulic head, causing an offshore flow of bottom resuspended, it either resettles or is transported in accordance
water, down the pressure gradient (Figs. 22, 23). This offshore with any prevailing currents. This process also applies to the
flow creates a shear stress on nearshore sediments, which com- enhancement of tidal-current-related sediment transport by
bines with the oscillatory shear created by waves to entrain storms. Bottom friction associated with both tidal and storm
FACIES MODELS REVISITED: CLASTIC SHELVES 351

Sediment Supply < Rate of Accommodation Creation: TRANSGRESSION

n
ce sio
rf a ro
S un e E
ce nt

ri
rf a me

Ma
S uvi n e
Ra

Shelf Floor Slope-basin floor


Disequilibrium, erosional surface(s) Gravity transport
ACCOMMODATION-DOMINATED Irregular grain size patterns
Shoreface bypassing
Sediment starvation
and condensation
SHELF Wave/storm/tidal reworking Ocean currents

FIG. 13.—A) An accommodation-dominated shelf has more available accommodation than sediment input, and is consequently
retrogradational or transgressive. This is the case for most modern shelves, at the close of the Holocene transgression. Embayed
coasts and estuaries (see Boyd et al., this volume) and transgressive shorelines are common. Most sediment on the shelf is
autochthonous, being left behind by the transgression. These are either relict or palimpsest if reworked into a new equilibrium
with shelf waves and currents. Ravinement and marine erosion surfaces can cover much of the shelf area, and abundant sand-
ridge fields occur in the area of preexisting coarse-grained sediments (redrawn from Swift and Thorne, 1991). B) An idealized
profile and summary of processes and features of an accommodation-dominated shelf (redrawn from Johnson and Baldwin,
1996).

processes results in greater sediment resuspension and entrain- minerals and associated fecal pellets in sediments at the bound-
ment, resulting in greater net sediment transport in the net ary of oxidizing seawater and reducing interstitial waters, typi-
direction of the combined flow. cally at depths greater than 60 m (Odin and Matter, 1981). Due to
various combinations of authigenesis or reworking, glauconite
Autochthonous Sediments can exist in almost all shallow-marine environments (Amorosi,
1995). An abundance of glauconite implies a combination of
Authigenic minerals can form in shelf sediments and be a continual reworking of the sediment and slow rates of clastic
major aid in identification of depositional environment. Glauco- input, but the mineral can be redeposited elsewhere as a detrital
nite is often a significant component of sediments that have been component. Kitamura (1998) found glauconite distribution to be
interpreted as shelf deposits. Glauconite occurs as part of the an unreliable guide to environment and sequence stratigraphic
sand fraction, and is generally thought to form from micaceous interpretation. Stonecipher (1999), in a study of authigenic min-
352 JOHN R. SUTER

Mobile
Mobile
Bay
Bay

Mississippi
Mississippi
River
River
Atchafalaya
Atchafalaya
Chenier
Chenier Plain
Plain River
River New
New
Orleans
Orleans
Estuarine
Estuarine
Plume
Plume

Atchafalaya
Atchafalaya Balize
Balize
Complex
Complex Complex
Complex

Deltaic
Deltaic
Plumes
Plumes

FIG. 14.—NASA MODIS satellite image , 1 March 2001, of the Mississippi delta and surrounding areas, northern Gulf of Mexico.
Buoyant or hypopycnal sediment plumes issue from the deltaic distributaries in both the Balize and Atchafalaya complexes, as
well as from several estuaries (e.g., Mobile Bay) along the coast.

Estuary Wind-Driven Along-Shelf Flows

ve s
ach Wa
Fro

s arine ear
Be av
e Estu e s a nd
Sh
s
nt

Rip W Plum Wa
ves
ity
pth rav
De ) Infrag
(m
Ver
t
Diff ical 10 rms
usio Internal Waves y Sto
n
o nb
20 nsi t ion
Physical and
su spe lida
Biologic Mixing Botto
m Bo Re o nso
of Sediment Column Layeundary n dC
30 na
sitiotion
r
p o
Steep Gradients De turba
In Cross-Shelf Sediment Flux 40 Bio

Inner Shelf Mid Shelf


FIG. 15.—Block diagram illustrating the major physical processes influencing sediment transport and deposition on clastic shelves
(redrawn from Nittrouer and Wright, 1994.)
FACIES MODELS REVISITED: CLASTIC SHELVES 353

Inflow from Continental Shelf


river mouth IIa

I IVa
V
IIb III

IVb

Offshore

FIG. 16.—Patterns and processes of sediment dispersal in different parts of the shelf (redrawn from Bentley, 2003; after Nittrouer and
Wright, 1995). Stage I, bedload deposition and channel-mouth-bar formation; Stage IIa, seaward transport as a hypopycnal
plume, with some flocculation and suspension settling; Stage IIb, seaward transport in hyperpycnal gravity flows; Stage III,
temporary deposition on shelf; Stage IVa, resuspension and transport in water column as suspended load; Stage IVb,
resuspension and transport in gravity-driven flow; Stage V, long-term accumulation.

eral distribution in the Shannon Sandstone, noted the possibility (Swift and Parsons, 1999). As with glauconite and phosphates,
of a “glaucony” facies composed of true glauconite minerals and shelf deposits may well contain siderite clasts and layers, but
a “verdine” facies, which includes other green-pelleted minerals that cannot be used as evidence of formation on a continental
(chamosite, an iron-rich chlorite, and berthierine, an iron-rich shelf.
kaolinite). The verdine facies, in contrast, is found on the inner Autochthonous organic matter in continental shelves is cre-
shelf and is “particularly abundant in estuarine facies or immedi- ated by phytoplankton from inorganic elements via photosyn-
ately offshore from fluvial deltas”. Chafetz and Reid (2000) thesis. The process may continue up to depths of as much as 100
describe cross-bedded “glaucarenites” or sandstones in which m but is usually limited to shallower depths in turbid waters.
glauconitic minerals constitute more than 25% of the composi- Owing to the increased supply of minerals and other nutrients,
tion, from Cambro-Ordovician strata in Texas and New Mexico primary productivity is usually enhanced in areas of riverine
(USA). These strata were interpreted as nearshore and tidal plumes or upwelling. Much of the organic carbon produced by
deposits, with glauconite being rapidly deposited due to differ- photosynthesis is mineralized by animals into shells or concen-
ent sea-water chemistry in the Cambro-Ordovician oceans from trated into fecal pellets. These various forms settle through the
that of the present day. We may conclude that continental shelf water column into the bottom sediments. Once there, organic
sands are likely to contain glauconite, but the presence of glauco- carbon may be consumed by infauna, destroyed by bacterial
nite in a sedimentary deposit does not constitute proof of a shelf decay, resuspended by waves or currents, and/or oxidized. If
genesis. organic carbon is supplied at a greater rate than the rate of its
Authigenic phosphates can form in areas of slow detrital destruction, enough may be preserved to become hydrocarbon
sedimentation along outer continental shelves, slopes, and abys- source material (Fig. 25).
sal plains. Modern phosphogenesis on the western margin of Preservation of organic material in shelf sediments requires
southern Africa occurs in areas of upwelling where phosphate some protection from oxidative processes. Depending on rela-
released by microbial degradation of organic matter during early tive concentrations of dissolved oxygen, seawater may be oxic,
diagenesis results in formation of phosphate nodules (Compton dysoxic, suboxic, or anoxic (Tyson and Pearson, 1991). Modern
et al., 2004). Phosphorite nodules, crusts, and/or grains can shelves are largely nonstratified, oxidative environments, but
become concentrated as lags in shelf sediments, but, as with oxygen-depleted water masses are known to exist in many
glauconite, the presence of phosphatic sediments can only be an areas, particularly in association with rivers (e.g., Boesch and
indicator, not proof, of a shelf origin of a given deposit. Rabalais, 1991; Rabalais et al. 1991, van der Zwaan and Jorrisen,
Siderite is also commonly found in marginal marine deposits. 1991; Rabalais et al., 2001). Anoxic conditions can develop
Its formation is variably considered the result of soil develop- because of stratification and isolation of bottom waters, extrac-
ment (Sullivan et al., 1997) or as an early diagenetic product in tion of oxygen by organisms, or decomposition of particulate
estuarine or marine sediments. Hansley and Whitney (1990) and dissolved allochthonous organic matter (Boesch and
concluded that siderite is most likely to form in swampy, mar- Rabalais, 1991). In marked contrast with the present day, many
ginal marine settings. black shales in the ancient record are interpreted to have been
Its occurrence in the Shannon Sandstone has been used to deposited under anoxic bottom water conditions, although this
support interpretations ranging from incised-valley deposits is increasingly debated (see papers and references in Harris,
(Sullivan et al., 1997) to forced regressive shorelines (Bergman, 2005). Many of these are highly important as hydrocarbon
1994; Walker and Bergman, 1993), to open marine sand ridges source rocks, and many were deposited in shelf settings in
354 JOHN R. SUTER

Palynofloral Assemblage
Distribution Diagram

Dinoflagellates

Mangrove

Palms

Ferns

B Forest

FIG. 17.—A) Muddy littoral drift system of the combined Amazon–Orinoco Rivers causes deposition and massive progradation
of “fluid” mudflats along the northeastern coast of south America (from Aslan et al., 2003). B) LANDSAT image (bands 2, 4;
7 August 1996) of the suspended-sediment plume of the Orinoco. A hypothetical palynological characterization of the plume
is shown, based on data from Muller (1959). Greens and yellows (pie diagrams) represent terrestrially derived spores and
pollen, whereas reds and pinks represent mangrove and dinoflagellates, respectively (degree of marine influence). In this
example, the fluvial discharge and associated “fresher” water dilute the marine palynofloral signature such that dinoflagel-
lates may be present only beyond the freshwater sediment plume. Mangrove pollen will be present in reduced numbers, and
the palynofloral assemblages will be dominated by terrestrially derived spores and pollen (image from Bureau of Economic
Geology website, palynological characterization by Thomas Demchuk, from Suter , 2001).

epeiric seas (Tyson and Pearson, 1991). Notable examples of peated sea-level fluctuations at various frequencies during Qua-
such source rocks include the Jurassic Kimmeridge Clay of the ternary glaciation, with consequent regression and transgression
North Sea (e.g., Miller, 1990) and various Mesozoic deposits in (Figs. 10, 11). Forced regression, or the natural basinward move-
North America (e.g., Nixon, 1973). ment of the shoreline during falling baselevel, provides an excel-
lent mechanism to emplace gravel and sand across a continental
Relative Sea-Level Change platform (e.g., Plint, 1988; Posamentier et al., 1992; Plint and
Nummedal, 2000). Ambulatory shorelines resulting from eustatic
Sea-level change is a major process affecting continental shelves fluctuations have been frequently invoked as precursors for
(Figs. 7–11). Modern continental shelves have experienced re- “isolated” marine sands in the subsurface and outcrop (e.g.,
FACIES MODELS REVISITED: CLASTIC SHELVES 355

Foreshore Backshore

Coa
Shoreface sta l Pla
Shelf ank in Che
Mud B niers
uds
ank M
Interb 0
ank
Mud B
m 10
10 k
20

30 m
Direction of ud
prevailing winds, c e ne M
o
Hol ud
waves, and currents
c e ne M
sto
Plei
Semi-fluid or fluid mud
Transgressive Lag:
Consolidated Mud
Shells and Sand
Sequence Boundary Sand
Vegetated Mud flats

FIG. 18.—Depositional model for accumulation of fluid muds in the shoreface and inner continental shelf (redrawn from Rine and
Ginsburg, 1985). Fluid mud is supplied to the nearshore zone by deltaic and estuarine sources in the form of hypopycnal plumes
(Fig. 14). Such plumes may reach well out onto the shelf or across the shelf break, depending on their own characteristics, slope
of the shelf, and the prevailing shelf winds, waves, and currents. Sediments are deposited as the plume loses momentum into the
receiving basin. Individual grains settle out as turbulence decreases, abetted by the process of flocculation. The overall deposit
shows oblique to alongshore progradational architecture. Excellent examples of these deposits occur along the coast of northern
South America (on which this diagram was based), sourced largely from the Amazon and Orinoco Rivers, and the western
coastline of Louisiana, USA, the Chenier Plain (see Fig. 19).

FIG. 19.—The Louisiana, USA, Chenier Plain has been built by sediments derived from the Mississippi River (Fig. 14), chiefly during
periods in which the main course of the river is located in the western portion of its delta plain. The Mississippi, currently in an
eastern position, is presently attempting to avulse into the course of the Atchafalaya River (Fig. 14), providing a renewed sediment
supply to the chenier plain. This map, redrawn from Neill and Allison (2005), depicts the facies of recent sediments accumulating
on the continental shelf offshore of the Atchafalaya River.
356 JOHN R. SUTER

Gravity
Self-generated Turbulence FIG. 20.—Types of hyperpycnal flows now
recognized from river-mouth dis-
charges. A) Autosuspended hyper-
Lessening gradient pycnal plumes, with suspension pro-
Frictional Drag duced by turbulence within the flow—
i.e., a “normal” turbidity current. Grav-
ity and turbulence maintain the flow
Minimum angle 0.7° until frictional drag or a decreasing
gradient result in deposition. These are
A. Autosuspending Gravity Flow believed to be relatively rare on conti-
nental shelves because relatively steep
gradients are required to produce and
maintain the flow. B) Wave–current
enhanced gravity flow, in which the
turbulence associated with waves and/
Gravity Lessening gradient or currents, abundant sediment sup-
Wave-current Turbulence Lower waves/currents ply, and a gradient above 0.03 degrees
Frictional Drag can produce a gravity flow, creating
downslope transport and broad distri-
bution of sediments across a shelf.
Minimum angle 0.03° Deposition results when frictional drag,
lowered gradient, and/or decreasing
wave–current turbulence decelerate the
B. Wave/current–enhanced Gravity Flow flow (redrawn from Bentley, 2003).

l Hurricane
Hurricane
pica Katrina
Katrina
o
tratr nt
Ex Fro

Hurricane Lili

A B
FIG. 21.—The northwest Gulf of Mexico is microtidal, and fair-weather wave energy is relatively low. However, the area is frequently
impacted by extratropical cyclones and hurricanes. A) Image shows a relatively unusual event of a hurricane and extratropical
storm impacting the area at the same time. Hurricane Lili hit the Louisiana coast early morning of 3 October 2002. Hurricane Lili
was a relatively weak Category 2 storm, unlike B) the much larger and more powerful Hurricane Katrina, which made landfall
in the central Gulf of Mexico shoreline on August of 2005. Storms like these are relatively rare, although they have been
increasingly common in recent years. On average, a hurricane impacts a particular area of the Gulf about every three years, but
occasionally multiple impacts occur, as in 2004 in the eastern Gulf of Mexico, and 2005 in the north-central Gulf. These images
illustrate the counterclockwise circulation of tropical storms in the northern hemisphere. Conservation of angular momentum
dictates that the northeastern quadrant of the storm is the most powerful, with the highest winds and highest storm surge. About
25–40 times a year, extratropical storms (known as “northers”) come through, substantially increasing the wave and current
energy on the shelf. These extratropical storms happen often enough to essentially be considered part of the normal process
environment. A. Image courtesy of Accuweather.com. B. NASA GOES-12 imagery over AMSR-E sea-surface temperature.
FACIES MODELS REVISITED: CLASTIC SHELVES 357

Storm Wave, Wind Setup

Set-up
Currents
MSL
Reference
horizon Pressure;
Waves Flood
Tide

Oscillatory boundary layer


A

HIGH LOW
PRESSURE GRADIENT
SHORELINE

CORIOLIS PRESSURE

FIG. 22.—Geostrophic flow is driven by the response of the water column to storms, in combination with the Coriolis force,
illustrated here for an idealized northern hemisphere shoreline. A) Surface waters are driven onshore by wind shear, wave
setup and shoaling, and storm surge (the so-called inverse barometer effect). B) The resulting pressure gradient drives bottom
water seaward, but the flow is deflected to the right (northern hemisphere) by the Coriolis force caused by the earth’s rotation.
Given sufficient time a balance between the two forces is achieved, and a geostrophic flow is developed parallel to bathymetric
contours (redrawn from Duke, 1990; Walker and Plint, 1992).

"Enhanced" Sea Level


Coastal Setup
"Normal" Sea Level

(s)
L INE Wind, wave setup;
E
OR pressure effects
SH Do wnw increase local sea level
e llin
g
Low-angle
Laminae
SH
Trough Cross Beds OR
EF
AC Evolves into
E Geostrophic Flow
Hummocky Beds
SHELF
Storm-Graded Sand Beds
Storm-Graded / Bioturbated Muds

FIG. 23.—Shoreface–shelf sediment succession formed in response to storm impact. Rip currents from the storm-enhanced surf zone
along with downwelling storm currents resulting from coastal setup carry sediment offshore onto the shelf. Interaction with the
Coriolis force leads to the evolution of geostrophic currents, sweeping storm-graded graded sediments obliquely offshore and
alongshore (redrawn from Swift et al., 1991).
358 JOHN R. SUTER

Wave
Shelf Shoreface base
Foreshore

Average
storms
Wave base of Shoreface
average storms

storms
Major
Wave base of Shelf
major storms

Grain size
Bed thickness
Amalgamation
Tempestite
frequency
Lamination

Bioturbation
Parauthocthonous Mixed Shell layers
FIG. 24.—Idealized vertical succession and summary of onshore–offshore trends from the Helgoland Bight of the southern North Sea
(redrawn from Aigner and Reineck, 1982).

Terrestrial
Terrestrial Flux:
Flux: Primary
Primary Productivity
Productivity
Organic
Organic matter,
matter, nutrients
nutrients

Organic
Organic Carbon
Carbon Flux
Flux

Water
Water
Column
Column
Mineralization
Mineralization Oxidation
Oxidation

Carbon in Bottom Sediments

Infauna
Infauna Resuspension
Resuspension
Burial
Burial

Total
Total Organic
Organic Carbon
Carbon
FIG. 25.—General scheme for the production of marine source rocks (redrawn from Schwarzkopf, 1993). Terrestrial flux (runoff,
hypopycnal plumes, hyperpycnal flows, storm deposits, atmospheric fallout) contributes particulate organic matter directly to
the shelf, as well as nutrients to stimulate primary productivity. These two contribute to flux of organic carbon, which settles down
through the water column. Under ordinary circumstances, over 90% of the flux of organic carbon on shelves is mineralized in the
water column and sediments (de Haas et al., 2002). Once in the bottom sediments, organic carbon may be consumed by infauna,
destroyed by bacterial decay, resuspended by waves or currents, and/or oxidized. If organic carbon is supplied at a greater rate
than the rate of its destruction, enough may be preserved to become hydrocarbon source material.
FACIES MODELS REVISITED: CLASTIC SHELVES 359

Walker and Plint, 1992; Bergman, 1994; Bergman and Walker, Sometimes overlooked in discussions of relative sea-level
1999). Transgression, both eustatic and autocyclic, and subse- fluctuations are their effects on fine-grained sedimentation.
quent reworking of preexisting shoreline deposits has been a When base level is positioned close to a shelf margin, silts and
favorite mechanism for the formation of shelf sands in Quater- clays may be delivered directly to slope and basin environ-
nary settings (e.g., Stubblefield et al., 1984; Penland et al., 1986; ments, bypassing the exposed shelf. Changes in base level alter
Penland et al., 1988), or at least for the provision of sediments that the size and distribution of microfossil and palynofloral zones
ultimately become shelf deposits. (Fig. 26). Familiar paleobathymetric zonations, such as inner
In extending Quaternary analogs to the ancient record, it is neritic, outer neritic, and bathyal, are most useful in transgres-
important to remember (1) the relatively unusual nature of the sive and highstand configurations like that of the present day.
late Quaternary—i.e., a time of high-frequency, high-magni- With base-level fall, the areas of these zones become greatly
tude base-level fluctuations, and (2) the relatively short-lived compressed and the terminology becomes less useful and more
nature of the base-level cycles. Quaternary analogs involving difficult to apply.
sea-level changes are best applied to those portions of the
Tertiary, for which the existence of substantial Antarctic ice SHELF PROCESSES AND DEPOSITS
can be documented, or the late Paleozoic, for which similar
sea-level cycles are interpreted. Late Paleozoic times present a The dynamics of clastic shelves can be exceedingly complex,
different plate-tectonic configuration than the present day, owing to the myriad of interacting processes (Fig. 6). Typically,
with fewer trailing margins and greater distribution of epeiric the most important of these processes are tides, waves, storms,
seaways. and semipermanent oceanic currents. One or the other of these

FIG. 26.—Vegetative and palynological response to base-level fluctuations. A) During conditions of a relatively high base level in a
transgressive and highstand setting, the influence of marine waters encroaches a significant distance landward, increasing the
suitable habitat for salt-water-tolerant species, predominantly mangrove. Thus, mangrove pollen as well as marine dinoflagel-
lates should be abundant and characteristic of these systems tracts. Terrestrial spore and pollen habitat is pushed back well into
the hinterland and may constitute a small percentage of the overall palynofloral assemblage. Fluvial drainage can take these
terrestrial palynoflora out onto the extensive continental shelves, mixing with the marine indicators mentioned above as well as
marine microfossils. B) In an idealized lowstand setting, opportunistic species of the savanna and hinterland occupy the exposed
shelf. Accordingly, the terrestrial pollen/spore habitat is abundant and their palynoflora can be transported significant distances
offshore, giving a more terrestrial signal. Mangrove areas are less extensive, and marine dinoflagellates are rarer owing to reduced
habitat. (Redrawn from Demchuk et al., 2004; based therein on Poumot, 1989).
360 JOHN R. SUTER

Storms
1 Bay of Fundy
2 North Sea
3 SE Africa

Wind-driven
4 East Coast North America

Thermohaline
4 6 5 Eastern Australia
6 NW Gulf of Mexico

3
c Cu
i rr e
ean nt
2 Oc s
5
Fair-weather processes
1
Tides Waves
FIG. 27.—Qualitative process classification of continental shelves, depending on the relative balance of “fair-weather” processes
(oceanic currents, tides, and waves) with storms (redrawn from Johnson and Baldwin, 1996). Numbers refer to shelf regions
discussed in this review. The Bay of Fundy of Nova Scotia, Canada, is included as the type section of tide dominance.

usually dominates a given shelf, leading to a process classifica- by differences in water temperature and salinity, and are associ-
tion (e.g., Johnson and Baldwin, 1996) similar to that for deltas ated with the sinking of dense, colder water at high latitudes.
(e.g., Galloway, 1975) and estuaries (Dalrymple et al., 1992) (Fig. Thus, the currents driven by thermohaline forcing are typically
27). The dominant processes on a given shelf area act upon found in deeper waters and are not major influences on sediment
previously existing deposits and/or sediments contributed di- transport on continental shelves. Wind-forced currents result
rectly from riverine or alongshore sources to produce a particu- from the frictional shear stress created by wind blowing over the
lar distribution of shelf deposits. It has been suggested that water surface, with energy transferred through the water column
about 80% of the world’s shelves are dominated by storm by turbulence. The surface circulation of the world ocean is
waves, 17% by tidal currents, and 3% by ocean–current interac- mostly wind driven. Current directions and resulting sediment
tions (Walker, 1984; Swift et al., 1986), but such classifications transport typically deviate from the originating wind directions,
are typically qualitative. Porter-Smith et al. (2004) published a owing to the apparent Coriolis force.
more quantitative classification of the Australian continental Major oceanic basins with bordering continental masses de-
shelves, using estimates of significant wave heights and periods velop large, essentially closed circulation patterns known as
and of tidal-current speeds. Based on their calculations, tidal gyres. These usually display east–west aligned currents at the top
currents dominate 17.4% of the shelf area, currents derived from and bottom of the system, and north–south boundary currents
tropical cyclones dominate 53.8% of the shelf area, ocean swell along the sides. These currents are effectively permanent, and are
and storm-generated currents dominate 28.2%, and intruding very important in the global heat balance. Many of the western
ocean currents account for about 0.6% of sediment transport boundary currents—e.g., the Gulf Stream in the North Atlantic,
(their Figure 1). Although the quantitative approach is encour- the Kuroshio in the North Pacific, and the Agulhas Current in the
aging, application to the stratigraphic record is still somewhat Indian Ocean (Fig. 28), can extend for as much as 1000 m through
problematic, inasmuch as dominant processes often vary sea- the water column and have flow velocities well in excess of that
sonally, and one process or another may dominate as the system needed for sand movement. Consequently, when these currents
evolves through tectonic or eustatic accommodation cycles intrude onto the continental shelf they can be very effective
(Leeder, 1999). sediment transport mechanisms, or, in the absence of sand, very
erosive.
Oceanic Currents However, such situations are comparatively uncommon.
Probably the best-known example of a current-dominated
Oceanic currents (Fig. 28) are typically classified according to shelf is in the southwestern Indian Ocean, where the Agulhas
their formative mechanism as either wind driven or thermoha- Current (Fig. 29) creates persistent southward flow across the
line (Fig. 6), or by the water depths in which they occur (surface, outer shelf offshore of southeastern Africa (Flemming 1980,
intermediate, deep, or bottom). Thermohaline currents are driven 1981; Ramsay, 1994). The current usually flows just offshore of
FACIES MODELS REVISITED: CLASTIC SHELVES 361

n
s t nd ia
E a nla eg
ee w

Lab
Gr or
N

r ado
Oyashio
Alaska

r
North Pacific North Atlantic
f Drift Pa c ific
ul North
G eam
tr
S Canary
California
North Equatorial
Kuroshio
Nort
Equatorial Equatoh
North Equatorial Counter rial North Equatorial
Equatorial South Equatorial Equatorial Counter
Counter
South Equatorial South
Equatorial
West
Australia East
Peru Brazil Australia
Benguela
Mozambique

n dian
tlantic th I
South Pacific South A Sou
Antarctic Circumpolar Antarctic Circumpolar
polar Subpolar
tic Sub Antarctic
Antarc

FIG. 28.—The major semipermanent ocean surface currents of the world. Many of these currents affect the continental shelves of the
world. Some, such as the Agulhas Current on the southeast Africa margin, shown in red, regularly impinge onto the continental
shelf. Compiled and redrawn from numerous sources.

the shelf margin, but it periodically intrudes onto the shelf covered in greater detail elsewhere in this volume (see Boyd et
(Fig. 30), with velocities up to over to 3 m/s (Ramsay, 1994), al.). Briefly, tidal currents, like tides themselves, may be of the
eroding fine-grained sands from nearshore deposits which are semidiurnal, diurnal, or mixed type. Where flows are restricted
swept seaward and transported downcurrent. A variety of to certain channels, as in rivers or estuaries, tidal currents are
bedforms, including huge dunes up to 17 m high, 4 km long, reversing, i.e., flowing alternately in approximately opposite
and over 1 km wide, are developed along the current pathway directions with a period of little or no current, called slack water,
(Flemming, 1981). The southeastern Africa shelf is character- at each reversal of the cycle. This produces the characteristic
ized by a number of headward-eroding submarine canyons rhythmic signature of coarse–fine, thicker–thinner laminites
that are continually supplied with sand by this process (Ramsay, used to recognize tidal deposits in the stratigraphic record (Nio
1994). Flemming (1980, 1981) indicated that most of the sand and Yang, 1991).
mined from the nearshore by the Agulhas current was trans- Tides in open shelf settings are not constricted into specific
ported off the shelf. channels, and are strongly affected by the Coriolis force (Figure
Some other areas of accumulation of outer-shelf sand bodies 31). As such, shelf tides rotate, varying continually through tidal
deposited by, or at least significantly influenced by, oceanic cycles. Although current strength does fluctuate periodically,
currents include the northeastern margin of Brazil (Vianna et al., rotary tides do not experience slack water periods, which has
1991; Testa and Bosence, 1999), the Campos margin of Brazil significant implications for the recognition of shelf tidal depos-
(Viana et al., 1998), the Grand Banks off of Nova Scotia (Dalrymple its. The rhythmic deposition of individual and double slack-
et al., 1992), and the Pennell Coast margin of Antarctica (Rodriguez water mud drapes that are cited as the best recognition criteria
and Anderson, 2004). In most cases, the coarse-grained sediments for tidal deposits (Nio and Yang, 1991) are unlikely to form
available to these currents are relict, and their deposits are thus under rotary tidal conditions. Mud deposition is, however,
palimpsest. quite possible, even without slack water. If suspended-sedi-
ment concentrations are sufficiently high (> 100 mg/l), mud is
Tides deposited, particularly if aided by flocculation or biogenic
processes (Galloway and Hobday, 1996). Still, mud drapes are
Tidal currents are generated by the differential forces be- extremely common features of many different deposits—peri-
tween the gravitational attraction of the earth–sun and earth– odicity and rhythmic variations in lithology and bed thickness
moon systems, combined with the centripetal acceleration pro- are the keys to the recognition of tidal deposits.
duced by the revolution of the earth around the sun and the Tidal currents can be quite significant in shelf settings, ca-
moon around the earth. These mechanisms, the resultant cur- pable of extensive bedload transport and formation of significant
rents, and their effects on estuaries and shorelines systems are sand bodies (Figs. 32–35). Tidal sand sheets and ridges cover
362 JOHN R. SUTER

0 800
A

FIG. 29.—Oceanic currents along the Southeast African shelf. A) The Agulhas current forms from the recombination of the
Mozambique and Madagascar currents, and flows along the South African shelf at velocities up to 3 m/s (Ramsay, 1994) (redrawn
from Johnson and Baldwin, after Flemming, 1981). B) The Sea-viewing Wide Field of view Sensor (SeaWiFS) image 28 March 1999,
shows the Agulhas Current being deflected southward along promontories on the shoreline, resulting in cross-shelf currents and
sediment transport. Higher chlorophyll concentrations are shown in red and orange (NASA Visible Earth SeaWiFS). Note the very
high productivity in the Benguela upwelling zone to the northwest.

thousands of square kilometers in tide-dominated shallow-ma- features are typically oriented subparallel to prevailing tidal
rine settings. Transgressive tidal sand sheets are typically only a currents and reach lengths of up to 180 km, widths of up to 10 km,
few meters thick (Dalrymple, 1992), but examples of much larger and thicknesses of up to 40 m. High-resolution seismic lines
sand ridges or banks abound on modern continental shelves, document complex internal stratigraphies and multi-stage histo-
including the North Sea (Figs. 31–33; e.g., Houbolt, 1968; Stride et ries (e.g., Fig. 35). Migration of these ridges is evident both from
al., 1982; Belderson et al., 1982; Belderson, 1986), Celtic Sea (Berné historical bathymetric changes and from packages of clinoform
et al. 1998; Reynaud et al. 1999); Georges Bank (Fig. 34; Twichell, reflections seen in seismic data. Dips of these lateral-accretion
1983; Dalrymple, 1992) and the East China Sea (Fig. 35; Yang and surfaces range from 5–10° in active ridges to less than 2° in
Sun, 1988; Yang, 1989; Liu et al. 1998; Berné et al., 2002). These “moribund” (inactive) ridges. In the East China Sea ridges shown
FACIES MODELS REVISITED: CLASTIC SHELVES 363

Coastline

Coastline
A B
Land Agulhas current

Wave-dominated
Current-dominated shelf
nearshore sands

Fluctuating boundary zone

FIG. 30.—Conceptual model for sedimentation on the outer shelf off South Africa, controlled by the Agulhas Current (redrawn
from Flemming, 1980). The current migrates laterally, moving A) onto and B) off of the shelf in response to seasonal
fluctuations. During the onshore phase, sand is eroded from the wave-dominated nearshore zone and transported down
current, and in some cases off the edge of the shelf, through submarine canyons and into deeper water (Ramsay, 1994). The
nearshore sand prism also migrates in response to the fluctuations of the current. The original model is not drawn to scale, but
the shelf off of southeast Africa is quite narrow, ranging from 3 to 20 km (Flemming, 1980, 1981; Ramsay, 1994).

in Figure 35, clinoformal reflectors dip obliquely to the long axis modern tide-dominated deltas and/or estuary settings (e.g.,
of the ridges and parallel to their steeper sides at angles of about Allen, 1991; Dalrymple et al., 2003), but these deposits would
2 degrees. Individual packages are about 1 km in width and about undergo substantial reworking upon transgression, and conver-
10 m thick. sion into tidal shelf ridges. Grain-size sampling and analyses,
How do the tidal ridges form? The most successful model to side-scan sonar images, and seismic profiling have shown some
date is that of Huthnance (1982), modified by Hulscher et al. of the surficial bedforms and internal structures of Quaternary
(1993). As geostrophic or tidal flows accelerate across a bathy- tidal sand ridges (e.g., Belderson et al., 1982; Stride et al., 1982;
metric irregularity on the sea floor, they tend to accelerate as Reynaud et al., 1999; Le Bot and Trentesaux, 2004; among many
they flow over the up-current side because of flow constriction others; see Figures 33, 35). Seismic profiles, side-scan sonar, and
and then weaken over the crest and down-current side (Fig. 36). multi-beam bathymetry show that active ridges in all settings are
This process perpetuates ridge growth and existence, as the up- covered by bedforms. Limited core information shows the pres-
current side is eroded and sand is deposited on the crest and lee ence of cross bedding (Dalrymple, 1992; Davis et al., 1993).
side, causing upward growth and/or down-current migration. Generally either flood or ebb currents dominate particular areas,
If sufficient sand is available, the process continues until an resulting in primarily unidirectional cross-bedding dips. Due to
equilibrium profile is reached, at which point upward growth diminishing current velocities, bed thickness decreases
ceases and widening of the feature begins. Solitary ridges form downcurrent, and the beds ultimately pass into rippled and
where there is only a single, initial perturbation (and limited burrowed fine sand and mud. Davis and Balson (1992) drilled a
sand supply), while fields of ridges develop where there are 35-m-thick, moribund tidal sand ridge in the East Bank complex
many nuclei and an abundance of sand. Calculated models for of the North Sea. Although recovery was limited and little infor-
flow across shelf sand ridges agree well with the observed mation was obtained about sedimentary structures, the overall
velocity field across Middlekerke Bank off the Belgian coast, deposit comprised two coarsening-upward successions of well-
although internally the ridge shows a multistage history (Fig. sorted fine sand, with bioturbation limited to the top meter of the
37; Berné et al., 1994). upper unit. The upward-coarsening trend is presumably the
The internal lithofacies and sedimentary structures of Quater- result of more intense current and wave action on the ridge crests
nary tidal sand ridges are not particularly well known. Some (Dalrymple, 1992). Davis et al. (1993) studied a series of smaller
facies models for tidal sand ridges have been proposed from tidal ridges off southwestern Florida. These ridges contain 3–4 m
364 JOHN R. SUTER

Time Tide Tidal Coriolis 10


NORWAY
(hours) levels 1m
flow influence
11 09
2m
00

0 H L
01 03

02
SCOTLAND
L
3

GERMANY
02
H
3m 03
01

04

00
6 L H
IRELAND
05

1m
06 11
10

09
2m
H 4m 07 08
ENGLAND 07 06
9
L WALES 03
11

00 02
3m 01 The NETHERLANDS

(initial
B T9
12 H L 4m BELGIUM
situation)
5m
FRANCE
A T0, 12 P
T6 C

T3

FIG. 31.—Coriolis effects cause tides on open continental shelves to take on rotary patterns. A) For a semidiurnal (twice-daily) tide,
a standing tidal wave in a basin with initially high water in the west (T0), sets up an eastward flow that is deflected to the right
(south) by the Coriolis effect, resulting in higher water on the south side at Time T3. The ensuing tidal flow to the north is then
deflected eastward, creating higher water on the eastern side of the basin at Time T6, six hours after the initial situation. The process
continues until the high part of the standing wave returns to its original position on the western side of the basin (T12). B) The net
effect is a rotary motion around the basin every 12 hours for a semidiurnal tide or every 24 hours for a diurnal tide. At the central
point “P” in Part B, there is no tidal motion (amphidromic point) and the whole rotary tide is called an amphidromic system. C)
This process is best known and illustrated in the North Sea, where the tidal range varies from 4 m on the English coast to nearly
zero on the coast of Norway. Solid lines are cotidal lines radiating from amphidromic points of zero tidal range, which indicate
the times of high water. Dashed lines show the increase in tidal range away from amphidromic points. Tides move south along
the eastern coast of Scotland and England, and high tide on the northeastern coast of Scotland occurs more than 12 hours before
high tide on the southeastern coast of England (redrawn from Clifton, 1999).

of clean, cross-bedded fine-grained quartz sand overlying muddy, Quaternary tidal sand ridges appear to evolve during con-
bioturbated sands. A subtle coarsening-upward trend character- tinuing transgression. Most present-day tidal ridge fields exist
izes the ridge sands, which contained variable amounts of coarse- because of tidal resonance, or enhancement of the open ocean
sand-size shell hash. tides, due to a particular basin configuration. This resonance
Berné et al. (1994) interpreted the Middelkerke bank of the may be a relatively brief feature in the history of a given locality,
North Sea (Fig. 37) as a composite feature, consisting of erosional and it be lost as water depths increase during transgression (e.g.,
remnants as well as combined tidal and storm deposits. Berné et Reynaud et al. 1999). In this case, as resonance is lost and tidal
al. (1998) made a similar interpretation for sand ridges in the energy decreases, and current speeds drop, causing sand ridges
Celtic Sea. These features, up to 60 m high, 200 km long, and 7 km to become wave-dominated and eventually moribund as the
wide (Berné et al. 1998), are morphologically similar to other tidal energy regime wanes. As the transgression proceeds, there is an
sand ridges but were interpreted as forming by erosional sculpt- upward change from cross bedding to finer-grained, rippled
ing of former lowstand deposits. However, Reynaud et al. (1999) and burrowed sediment as current speeds decrease, which
believe that the same ridges formed entirely in a shelf environ- should give rise to a fining-upward succession. Sediment is
ment, albeit under evolving conditions of tide and wave domi- eroded from the ridge crests by storm waves and deposited as
nance. The latter two studies are based on seismic reflection a drape on the ridge flanks, producing a laterally accreting
profiles and provided minimal lithologic information. deposit. Presumably a coarsening-upward profile may be pre-
FACIES MODELS REVISITED: CLASTIC SHELVES 365

FIG. 32.—A series of sand ridges align


subparallel to the tidal currents gen-
erated in the North Sea (redrawn
from Snedden and Dalrymple, 1999,
after van den Meene, 1994). The
tidal currents rework relict sands
England from the last glacioeustatic lowstand
The into large sand bodies of various
Netherlands dimensions.

Belgium
ns
bo
Rib
nd
Sa
e

gth
dg

FIG. 33.—General model for a tidal sand trans-


Ri

port pathway in the North Sea. Sediment


n
nd

Stre

is eroded from the seafloor at the up-


Sa

stream end, leaving a shelly gravel lag in


which large wave ripples may be present.
ent

Flow-parallel sand ribbons and isolated


dunes occur in the zone of bypassing,
ur r

Large where neither erosion nor deposition oc-


Dunes curs. The reworked sediment is trans-
gC

ported down current and deposited where


sin

tidal current speeds decrease. In deposi-


tional areas, sand sheets may cover thou-
rea

sands of square kilometers. Dunes of vari-


ous scales are extensively developed on
De c

these sheets. Sand ridges form if current


Small speeds are in excess of 1 m/s and suffi-
Dunes cient sand is available (redrawn from
Belderson et al., 1982).

Rippled
sand sheet
366 JOHN R. SUTER

~ 5 km

FIG. 34.—Digital multibeam bathymetric image of the Canadian portion of Georges Bank. Water depth in this image is about 50–70
m, and ridges are approximately 6–10 m in height; the greatest ridge relief measured in the survey was 18 m. Currents sweep
south-southeast from the Gulf of Maine across the Georges Bank in this area (about 50 to 70 m deep), providing the mechanism
for sediment transport. Sediments on the shelf are largely palimpsest, derived from preexisting glacial deposits. The tidal currents
are rotary with maximum velocities ranging from about 30 to 75 cm/s. Transient currents from wind-driven forcing, storms, and
eddies off the Gulf Stream also influence this area. (Image and data courtesy of Brian Todd, Geological Survey of Canada Atlantic
Geoscience Center, 2003.)

served if the change in tidal regime is rapid and current speeds eastern coast of South America (Johnson and Baldwin, 1996;
diminish so quickly that the ridge is effectively “frozen.” Greater Galloway and Hobday, 1996).
wave energy at the tops of the ridges may also contribute to an Given sufficient wave energy and resultant current strength,
upward-coarsening profile. Ultimately, the ridge may become very significant longshore sediment transport pathways can be
draped with shelf muds. established, as is the case for eastern Australia (Figs. 38–41;
Boyd et al., 2004a; Boyd et al., 2004b). The Eastern Australia
Wave- and Storm-Generated Processes current (Fig. 28) runs south along the continental margin, sweep-
ing the outer continental shelf of clastic sediments, confining
Unlike semipermanent oceanic currents and tides, wave and deposition to the nearshore. Inshore, a longshore-drift system
storm processes result from meteorological forcing. Clifton (this runs for over 1500 km along the coast. At the end of the sediment
volume) gives a good introduction to wave and surf-zone pro- transport pathway past Fraser Island, tidal currents issuing
cesses as they influence shoreface and shoreline environments. from Hervey Bay intercept the fine- to medium-grained sands
Waves shoal as they enter shallow water, moving both sediment brought in by longshore currents, creating a cross-shelf-migrat-
and water onshore. The onshore setup of water creates a hydrau- ing dune field (Figs. 39–41). The dunes, which are mantled by
lic head that drives rip currents and longshore currents. The smaller, superimposed bedforms, have reliefs up to 6 m and
characteristics of wind waves are dependent on three main wavelengths of 200–700 m. Internal structures of the dunes are
factors: the strength and duration of the wind, and the distance not yet available, but we can envision composite sets of large-
over which the wind blows (the fetch). Consequently, wave- scale cross-bedding as dominant sedimentary structures. This
dominated shelves tend to be those open to oceanic waves, such relatively unusual situation provides a mechanism for sand
as the Pacific Northwest of North America or the eastern coast of transport across the shelf, onto the continental slope, and into
Australia (Figs. 1, 28, 38). Porter-Smith et al. (2004) calculated that deeper water, serving to illustrate the power of combined flow
oceanic swell is able to entrain fine sand at depths of up to 142 m on continental shelves.
in portions of the Australian continental shelf. Other examples of Almost all shelves are affected by storms, including the famil-
wave-dominated shelves include the Bering Sea and the north- iar tropical cyclones (hurricanes and/or typhoons), mid-latitude
FACIES MODELS REVISITED: CLASTIC SHELVES 367

A Recent
Changjian
N
River Delta

Shanghai 0 100

km

East China Sea


Figure 37b
37B

Paleo-estuary
Mouth

B Direction of Ridge Migration


SW NE

20 m

00 km 5 5 km

FIG. 35.—Modern sand ridges in the East China Sea are interpreted to represent the retreat path of the Changjiang River Delta
during the Holocene transgression A) Individual ridges are 10–60 km long, 2–5 km wide, and up to 20 m in height. High-
resolution seismic lines B) document a complex internal stratigraphy and multi-stage history. Packages of clinoforms dip
obliquely to the long axis of the ridges, and parallel to their steeper sides at angles of about 2 degrees. Individual packages are
about 1 km in width and about 10 m thick. No paleocurrent data are available from the ridges, but the ridges are oriented
subparallel to the tidal currents in the East China Sea (redrawn from Posamentier, 2002; Yang and Sun, 1988; Liu et al., 1998;
Berné et al., 2002).

low-pressure systems, as well as high-latitude “disturbances” Tempestites have a fairly characteristic signature, making their
(Fig. 21). These systems engender a number of powerful pro- recognition in the rock record in core and outcrop comparatively
cesses on clastic shelves, including wind-driven currents, large straightforward. Identification of storm beds is one good way to
waves, coastal setup and downwelling, and, ultimately, geo- infer that a particular succession represents a shoreface–shelf
strophic currents (Figs. 21–23). Storms and resulting currents are environment, although distinguishing between the two usually
responsible for a variety of deposits on shelves, including requires further context (see Clifton, this volume). The general
tempestites, sand sheets, and sand ridges. succession of tempestites has been widely described (e.g., Aigner
and Reineck, 1982; Dott and Bourgeois, 1982; Aigner, 1985; Walker,
Tempestites 1984; Leckie and Krystinik, 1989) (Figs. 24, 42, 43). Waves,
downwelling storm currents, or some type of combined flow cut
Storm beds, or tempestites, occur widely on modern shelves an initial erosional surface. The surface may be flat, undulatory,
(e.g., Fig. 42) and are extensively interpreted in the ancient record. or channelized at a small scale (gutter casts) and may contain sole
368 JOHN R. SUTER

Coarsest
GRAIN SIZE
Finest

Highest
SHEAR STRESS
Lowest FIG. 36.—Model for the growth of a shelf sand ridge from
an initial topographic high. Accumulation of sand on
the high is due to a phase lag between topography and
bottom shear stress (Redrawn from Nummedal and
Suter, 2002; modified therein from Swift and Field,
1981, based on Smith, 1970).

0 NW SE 0 NW SE
500 1000m
A 500 1000m
Middelkerke Bank F
10
Middelkerke Bank 10
U7
U6
U7 U3
20 20 U4
U4
U3
30m Tertiary 30m Tertiary
0 NW SE
500 1000m
Middelkerke Bank B 0 NW SE
10
500 1000m
Middelkerke Bank G
U7 10 Oostende Bank
20 U4 U7 U6
U2 20

30m
U2 Tertiary U4
30m Tertiary
0 NW SE
500 1000m
Middelkerke Bank C 0 NW SE
10 500 1000m
H
U7 10 Middelkerke Bank
20
U6 U7
U4 U3 U6
U2 20
30m
Tertiary U1 U4
Tertiary
0 NW SE
500 1000m
Middelkerke Bank D 0 NW SE
10
Oostende Bank 500 1000m
I
U7 U6 10 Middelkerke Bank
20 U4 U6
U7
U2 20 U4
30m U1
Tertiary Tertiary
30m

0 NW 500 1000m
Middelkerke Bank
SE
E A 0 NW SE
Oostende Bank B
500 1000m

Middelkerke Bank
J
C
D 10
U5 E
F U6 U7
U3 G 20 U4
U2 H
I
U1 J Tertiary
30m

FIG. 37.—Interpreted high-resolution seismic profiles across the Middelkerke Bank, part of a field of tidal sand ridges offshore
Belgium in the North Sea. Note the multiple erosional surfaces within the ridge, and the upward change from channelized units
to more tabular packages. Vertical exaggeration is about 50:1; inclined reflections are actually dipping at less than 1 degree
(redrawn from Berné et al., 1994).
FACIES MODELS REVISITED: CLASTIC SHELVES 369

FIG. 38.—Merged satellite topographic-bathymetric image of the Southeast Australia continental margin (Boyd et al, 2004a). Wave
approach from the southeast sets up a strong northward-flowing longshore current and sediment transport system, which is
deflected across the shelf by tidal currents in the area of Fraser Island. Image courtesy of Ron Boyd, 2004.

marks or tool marks in the form of flute casts, groove casts, or sole have characteristics very similar to Bouma sequences (e.g.,
casts. Pebble or intraclast lags sometimes occur above the surface, Hamblin and Walker, 1979), and these were interpreted as the
which is overlain by low-angle-laminated to hummocky cross- deposits of shallow-water turbidity currents that took sediment
stratified sands (Harms et al., 1975). As storm strength wanes, below wave base.
wave oscillation ripples and/or combined-flow ripples form. The concept of shallow-water turbidity currents was ques-
Post-storm or normal background deposition is generally charac- tioned on several grounds. Firstly, no mechanism was available
terized by suspension fallout. for ignition of turbidity currents on the very low gradients that
Tempestites often have a characteristic ichnological signa- characterize the shelves of the Gulf of Mexico (<< 1 degree).
ture, which aids in their recognition (Figs. 44–46; Pemberton et al., Morton (1981), citing physical oceanographic measurements and
2001). Following the abatement of storm conditions, opportunis- the distribution of the storm bed, suggested that wind-forced
tic suspension-feeding faunas colonize the upper parts of the currents during the landfall of Hurricane Carla, rather than
sand and create dwelling structures of the Skolithos and proximal storm-surge ebb, were responsible for the deposition of the sand
Cruziana ichnofacies. Depending upon rates of sedimentation bed. Many authors (e.g., Aigner, 1985; Swift et al., 1986; Snedden
and storminess, an individual tempestite may be buried and et al., 1988, among others), presented evidence for geostrophic
preserved by fair-weather sedimentation, thoroughly churned flows as the dominant mechanism during storms, which would
by bioturbation into an unrecognizable mass, eroded away by tend to move sands obliquely offshore and ultimately along shelf
subsequent storms, and/or amalgamated with younger contours. This model is widely held today, although questions
tempestites (Fig. 46). A return to fair-weather conditions results persist. Snedden et al (1988) and Snedden and Nummedal (1991)
in deposition of finer-grained sediments with equilibrium mapped the Carla sand bed on the Texas shelf (Fig. 42) but did not
ichnological assemblages dominated by deposit-feeding and graz- follow the deposit into shallower depths, where the sand had
ing structures of the distal Cruziana and Zoophycos ichnofacies been thoroughly bioturbated. Leckie and Krystinik (1989) ob-
(Figs. 45, 46). Repetition of this process results in a cyclical served that most current indicators in storm deposits pointed
deposit, often containing a laminated to scrambled bedding style, offshore rather than alongshore, and questioned the importance
colloquially called “lam-scram” (Howard, 1975) (Figs. 46, 47). of geostrophic currents. Duke (1990) and Duke et al. (1991)
Early tempestite facies models invoked density currents to offered a solution—storms create combined flows, and instanta-
move sediments offshore. Hayes (1967) noted the deposition of a neous current indicators of the downwelling storm flows are
sand bed on the Texas shelf following the passage of Hurricane directed offshore, whilst net sand transport is obliquely along the
Carla in 1961 (Fig. 42). He theorized that following hurricane shelf, dominated by geostrophic flows.
landfall the storm surge ebbed back onto the shelf as density or The controversy regarding shallow-water gravity flows and
turbidity currents, flowing through washover channels cut into tempestite deposition has “reignited” (Myrow and Southard,
the shoreline by the storm surge. Even though subsequent at- 1996; Myrow et al., 2002)). New data and theories have been
tempts to find the Carla sand bed in the nearshore zone either expounded (e.g., Traykovski et al., 2000; Wright et al., 2001;
were unsuccessful, or discovered only thoroughly bioturbated Bentley, 2003) which argue that turbulence from storm waves
sand, the idea of a storm-surge turbidity current had impact. and/or currents creates enough suspended-sediment concentra-
Numerous ancient sandstones interpreted as storm deposits tion in bottom water layers to reach sufficient densities to activate
370 JOHN R. SUTER

A B rree
s
nntts
rr
CCuu
HERVEY
HERVEY

Tidal
Tidal
BAY
BAY ~~ 60km
60km

FRASER
FRASER
MAINLAND
MAINLAND ISLAND
ISLAND

Hervey FRASER
FRASER
Bay TIDAL
TIDAL ISLAND
ISLAND
DUNES
DUNES

RE NTTS
EN S
C
C UR
U RR
DAAL
L
TTIID

C
FIG. 39.—A) Map of Hervey Bay, Fraser Island, and Breaksea Spit, northeast Australia. The island and attached spit have prograded
to near the shelf margin, which turns to the northwest at this location (Fig. 38). B) The extensive longshore sediment transport
system seaward of Fraser Island and Breaksea Spit is intercepted by tidal currents moving in and out of Hervey Bay, resulting
in C) cross-shelf transport of sand in a series of large tidal dunes prograding to the shelf margin and beyond into deep water (Boyd
et al., 2004b). Map and photographs courtesy of Ron Boyd, 2004.

and sustain gravity flows (Fig. 16). The “wave-modified turbid- coarse silt and very fine to fine sand, but in larger grain sizes
ite” model (Fig. 48) suggests that combined flow is again the larger ripples and more steeply dipping strata occur (Leckie,
driver of shelf deposition, but in this case the combination is 1988). Definitive examples of hummocks on modern shelves are
wave-induced turbulence, geostrophic currents, and sediment still wanting, likely because their presumed formation during
gravity flows. This revival of shelf turbidites mirrors the increas- large storms precludes direct observation. Bouma et al. (1982)
ing emphasis being given to hyperpycnal density flows in under- reported HCS from box cores taken on the South Texas continen-
standing the delivery of sediments from rivers to ocean basins. tal shelf (their Figure 20). Swift et al. (1983) provide side-scan
Hummocky cross-stratification (HCS) (Fig. 43) has its own sonar records of what they term “hummocky megaripples,”
controversies. There is substantial agreement that storm waves which are likely candidates, although their formation was con-
are involved in the formation of HCS, but, beyond that, consensus cluded to be the result of combined geostrophic flows and storm
is lacking. HCS is widely used as an indicator of the lower- waves. Ramsay (1994) reported hummocks and swales in fine-
shoreface environment; however, given that storm waves can grained sand with trough–crest amplitudes of 40 cm and wave-
entrain fine sand sediment at depths of well over 100 m (Porter- lengths of 1.5–5 m at depths of 30–60 m from the shelf off of
Smith et al., 2004), these features can obviously occur in shelf southeast Africa. Dalrymple and Hoogendoorn (1997) interpreted
settings (Clifton, this volume). Original ideas on the formation of HCS and storm-graded beds within vibracores from shoreface-
HCS involved combined flows (e.g., Swift et al., 1983; Nøttvedt attached ridges off of Sable Island on the Scotian Shelf. Li and
and Kreisa, 1987). However, flume studies (e.g., Arnott and Amos (1999) observed symmetrical “large wave ripples” forming
Southard, 1990) indicated that HCS forms under almost purely at 39 m water depth during storms on the Scotian shelf. They
oscillatory conditions, with the addition of relatively minor uni- interpreted these features as combined-flow hummocky mega-
directional forcing causing a transition to low-angle laminae ripples. The low-angle laminae found in many vibracores from
dipping in the direction of current flow. HCS per se is limited to shelf deposits (e.g., Fig. 42) probably represent HCS.
FACIES MODELS REVISITED: CLASTIC SHELVES 371

Breaksea Hervey Bay


Spit
Tidal
Tidal Dunes
Dunes

Longshore
Longshore
Transport
Transport

N
N
5 km

FIG. 40.—Digital bathymetry from the area immediately offshore of Breaksea Spit, northeast Queensland, Australia (Boyd et al.,
2004b). Eastward-flowing ebb tidal currents from Hervey Bay to the west intercept and divert the massive northward-flowing
longshore transport system, forming a sand sheet of tidal bedforms progressing across the narrow shelf and onto the shelf margin.
Imagery courtesy of Ron Boyd, 2004.

Sand Bodies on Storm-Dominated Shelves are still in active process environments and are being trans-
formed by present shelf processes, although those in the deep-
Sand sheets and linear sand bodies are nearly ubiquitous est water seem to be moribund (Goff et al., 1999; Goff et al.,
on storm-dominated Quaternary continental shelves (Table 1) 2005).
and have been the subject of considerable research and publi- Many of the best studied—or in any event the most discussed
cation (see Table 1), with the most recent and comprehensive and debated—Quaternary storm-sand ridges occur on the conti-
review by Snedden and Dalrymple (1999). In subsurface and nental shelves of North America (Figs. 50–54). Sand ridges of the
outcrop interpretations, as well as on Quaternary shelves, Atlantic margin range from about 1 to 10 m in thickness and
storm-dominated ridges occur anywhere from just beyond the average around 3 km in width (Stubblefield et al., 1984; Goff et al.,
shoreface to more than 100 km from the “coeval” highstand 1999). Sand bodies of the Gulf coast are of similar dimensions.
shoreline. Shoreface-attached ridges are generally oblique to Mississippi Delta shoals are 4–6 m thick, 8–10 km wide, and up to
the coeval shoreline, whereas offshore ridges tend to be subpar- 50 km in length (Figs. 53, 54; Penland et al., 1989), and shoals along
allel to the coast (Fig. 49). Despite decades of research, there is the Mississippi–Alabama, western Louisiana, and eastern Texas
a great deal we do not know about these deposits. These modern shelves (Curray, 1960; Nelson and Bray, 1970; Frazier, 1974; Suter,
shoals and ridges are subparallel to the shoreline and can 1987; Suter et al., 1987; McBride et al., 1999; Rodriguez et al., 1999)
reliably be considered to be currently subaqueous and on the fit within a similar range. Vibracores through many of the Qua-
continental shelf, but the actual mode of their formation and ternary features (e.g., Penland et al., 1986; Swift et al., 1986; Rine
ultimate fate are still matters of debate and ongoing research et al., 1991, Hoogendoorn and Dalrymple, 1986; Snedden et al.,
(Figs. 50, 51). The general consensus today is that there are 1994, Dalrymple and Hoogendoorn, 1997; Snedden et al., 1999)
various types of continental-shelf sand bodies, the formation show the ridge lithofacies dominated by sand. Where not wholly
of which is related to transgression and probably involves the deformed by the coring process, these sands are dominated by
Huthnance mechanism (Berné et al., 1998; Dyer and Huntley, low-angle laminae, high-angle cross-bedding, graded storm beds,
1999; Snedden and Dalrymple, 1999). Many of these deposits hummocky cross-stratification, or bioturbation. Subordinate
372 JOHN R. SUTER

A. Bathymetric
Profile

B. Bathymetric Bar
Map Crest

Superimposed
Dunes

FIG. 41.—A) Bathymetric profile and B) digital multibeam bathymetry of tidal barforms crossing the continental shelf offshore of
Breaksea Spit (Boyd et al., 2004b). The large bars, with superimposed smaller dunes, are up to 6 m in relief, with wavelengths
of 200–700 m. A model of a Russian Oskar class submarine is shown for scale. Image courtesy of Ron Boyd, 2004.

amounts of mud and silt are present, but as individual beds reflections, landward-dipping reflections, as well as channeling
within distinct units, not distributed throughout the sand bodies. and other features within various units (e.g., Rodriguez et al.,
These may reflect preexisting deposits upon which the shoals 1999).
nucleated (cf. Stubblefield et al., 1984, Penland et al., 1988; Snedden
et al. 1984; 1999; Pope et al. 1989; Rodriguez et al., 1999; Snedden SHELF SANDS IN OUTCROP AND SUBSURFACE
and Dalrymple, 1999).
The basal contacts of Quaternary ridges are variable. In many Although many outcropping and subsurface sandstones have
cases, the sand bodies lie on composite erosional surfaces formed long been interpreted as shelf deposits (Tables 2, 3), these inter-
by a combination of wave ravinement and shelf currents. Others pretations remain controversial. “Isolated marine sand bodies”
have gradational basal contacts, probably reflecting the different (cf. Bergman and Snedden, 1999) like the Shannon and Sussex
origins of the basal portions of the sand bodies. Thus, the shelf Sandstones in Wyoming and the Tocito Sandstone in New Mexico
sand facies per se has an erosional lower boundary (Fig. 54), while are probably the best known and most studied of these inter-
some portions of the basal contact of the ridge may be grada- preted shelf sand ridges (Spearing, 1976; Brenner, 1978; Bouma et
tional. Internally, the contact between the shelf sand and its al., 1982; Tillman and Martinsen, 1984, 1987; Gaynor and Swift,
precursor is an erosional surface. 1988; Swift and Parsons, 1999; Nummedal and Riley, 1999).
Bathymetric map analyses, seismic profiling, and side-scan Despite this wealth of investigation, or perhaps because of it,
sonar imagery have shown that many shoals are active and alternative interpretations of these units abound: estuary-mouth
migratory. Side-scan sonar and multibeam bathymetric images shoals or tide-dominated bayhead deltas within incised valleys
show that bedforms migrate obliquely along the Atlantic ridges (e.g., Jennette and Jones, 1995; Sullivan et al., 1997), forced regres-
(e.g., Swift et al., 1986; Goff et al., 1999). Historic mapping and sive shorelines (Walker and Bergman, 1993; Bergman, 1994, 1999;
seismic profiling have documented migration of shoreface- Bergman and Walker, 1999).
attached ridges along shore (McBride and Moslow, 1991, Earlier interpretations of these deposits as shelf sandstones
Dalrymple and Hoogendoorn, 1997; Snedden et al., 1999). In the came largely from the occurrence of the sandstones within thick
Gulf of Mexico, landward migration of Ship Shoal of more than successions of demonstrably marine mudstones, as well as their
1.5 km in the last 100 years was documented by bathymetric reconstructed paleogeographic position well offshore of pre-
map analyses (Penland et al., 1988). Seismic reflection profiles of sumed coeval shorelines (e.g., Spearing, 1976; Tillman and
other shelf ridges in the Gulf of Mexico show seaward-dipping Martinsen, 1984, 1987). The Shannon Sandstone lies about 100 km
FACIES MODELS REVISITED: CLASTIC SHELVES 373

A C

FIG. 42.—The distribution and internal characteristics of the Hurricane Carla storm sand layer on the central Texas shelf. A) Net sand
thickness (cm) of the Carla storm bed as determined by Snedden and Nummedal (1991), extending obliquely across the shelf for
some 200 km (redrawn from Snedden and Nummedal, 1991). B) Graded sand layer found on the inner shelf by Hayes (1967)
immediately following the landfall of Hurricane Carla in 1961. Hayes (1967) interpreted deposition of the sand by storm-surge-
ebb turbidity currents, but Snedden and Nummedal (1991) followed Morton (1981) by arguing in favor of a geostrophic flow
mechanism (redrawn from Johnson and Baldwin, 1996). C) Close-up photograph of tempestites from offshore of Ship Shoal, a
sand body on the continental shelf off of the Mississippi Delta.

offshore of its presumably coeval shoreline, with a roughly shore- Vertical trends within the sandstones are inconsistent, rang-
parallel orientation reminiscent of the shelf sand ridges offshore ing from upward cleaning to blocky to upward fining. Depend-
of the Atlantic coast of the USA. The encasing marine mudstones ing on the perspective of the investigators, these have been
create ideal stratigraphic traps for hydrocarbons, and several of variably explained as the result of the original formation of the
these units comprise large reservoirs (Table 3). For example, sandstones as shoreface and/or incised-valley deposits or as part
Hartzog Draw (e.g., Tillman and Martinsen, 1987), which is of the normal succession of processes expected within a shelf
dominantly a Shannon Sandstone field, has an estimated 400 ridge. Thicknesses of the individual deposits are variable, but
MMBO in place. This economic driver engendered significant individual Shannon “ridges” approach 30 m in thickness, or
interest and research into Quaternary shelf sands, along with a several times that seen for Quaternary storm-dominated shelf
corresponding increase in shelf-sand interpretations in ancient sands.
rocks. As previously noted, one of the major problems with the
Summarizing characteristics of many units listed in Tables 2 shelf-sand interpretation was that of supply of coarse-grained
and 3, these deposits comprise sands, in some cases conglomer- sediment. No convincing mechanisms were available for the
atic, with varying amounts of glauconite and siderite. They are transport of sand and gravel tens of kilometers from coeval
commonly bioturbated, with ichnofacies ranging from the distal shorelines to be encased in muds far out on the shelf. Shelf
Cruziana to the proximal Skolithos. Primary sedimentary struc- plumes were often invoked, but such plumes comprise mostly
tures include hummocky cross-stratification, low-angle lamina- suspended material and are thus limited to finer-grained sedi-
tion, tabular and trough cross-bedding, sigmoidal cross-stratifi- ment (Scheihing and Gaynor, 1991). In addition, plumes are
cation, wave, current, and combined-flow ripples, mud drapes, generally turned shore parallel by geostrophic currents before
double mud drapes, elongate rip-up clasts, and rhythmites. Thick- reaching the offshore distances needed for some of the units in
nesses range from 10 to 30 m, widths from 5 to 10 km, and lengths question. Some of the “isolated marine sands,” such as the
of up to 160 km. Paleocurrents derived from cross bedding are Prairie Canyon Formation of Utah and Colorado (Cole and
generally unidirectional but are typically oblique to the sand- Young, 1991; Cole et al. 1997) do consist mostly of very fine
body orientation. sand, silt, and clay, trend subparallel to interpreted shorelines,
374 JOHN R. SUTER

Structures
Structures

1.1 Guttercast
Gutter Cast ๷
2.2 Solemarks
Sole Marks ๷
3.3 Partinglineation
Parting Lineation ๷ Perpendicular
๷ Perpendicular
4.4 Wave-ripplecrests
Wave-ripple crests ඳ Parallel
ඳ Parallel
5.5 Combined flowripples
Combined-flow ripples ๷
6.6 Currentripples
Current ripples ๷
FIG. 43.—Leckie and Krystinik (1989) compiled the directional and paleocurrent relationships of tempestites from a number of
Cretaceous sands. The orientations of wave-ripple crests (4) from the lower-shoreface to inner-shelf zones of wave-dominated
coasts are approximately parallel to the local paleoshoreline trend. Numerous current indicators, such as gutter casts (1), sole
marks (2) and parting lineations (3) from tempestites and sole marks from hummocky beds indicate that flows transporting
sediment offshore were at high angles to the paleoshoreline. Wave approach was also at high angles to the shoreline, as indicated
by the shore-normal orientations of sole marks, parting lineation, and asymmetric ripples on hummocky cross-stratification and
their orthogonal relationship with respect to orientation of wave-ripple crests. Compiled paleocurrent data indicate that HCS
forms under dominantly oscillatory flow with a weak, shore-normal, combined-flow component. This apparent conflict between
geostrophic current theory and observations from modern shelves was reconciled by Duke (1990) and Duke et al. (1991), in which
the offshore-directed indicators were attributed to instantaneous combined flow, with overall net transport still believed to be
dominated by geostrophic currents (redrawn from Leckie and Krystinik, 1989).

show an overall upward-cleaning vertical profile, and in gen- no real mysteries. The difficulty lay in how that sediment was
eral are more convincing as shelf plume deposits than coarser- reworked—was it reformed into shelf sands or did it retain its
grained units like the Shannon Sandstone. However, the Prai- original characteristics? In the latter case, was it originally depos-
rie Canyon has also been interpreted as incised-valley deposits ited as some type of shoreline, an incised valley, or some interme-
(Hampson et al., 1999); and similar units in the area have been diate deposit? All of these possibilities, and more, have been
interpreted as deposits of hyperpycnal flows (e.g., Pattison, proposed (see papers in Bergman and Snedden, 1999). Suter and
2005). Clifton (1999) examined the competing interpretations for the
The realization that Quaternary shelves had experienced Shannon Sandstone and found that none of them were able to
repeated glacioeustatic fluctuations solved the problem of sedi- explain all of the features found in the deposit. Nummedal and
ment supply problem (Figs. 10, 11). With ambulatory shorelines, Suter (2002) reviewed the formation of shelf sand ridges, and
repeatedly prograding and reworked by transgressions, coarse- followed Snedden and Dalrymple (1999) by concluding that
grained sediments on an otherwise fine-grained shelf presented most, if not all, were of composite origin.
FACIES MODELS REVISITED: CLASTIC SHELVES 375

Sandy Shore

Sandy
Backshore

Semi-consolidated
Substrate

Rocky Sublittoral Zo
Coast ne
Peat or
Xylic

B
at
Substrate

hy
al
Zo
ne
Abys
sal Zone

1 6 7 14 18
11 23
16
8 19 29
3 26
13 20
5
2 9 15
24
22
12 21 28 30
4 10 17
25
Peat or Xylic Semi-consolidated 27
Rocky Coast Substrate Substrate Sandy Backshore Sandy Shore Sublittoral Zone Bathyal Zone Abyssal Zone

Trypanites Teredolites Glossifungites Psilonichnus Skolithos Cruziana Zoophycos Nereites

FIG. 44.—Seilacher’s (1967) model of ichnofacies distribution, based on a number of environmental parameters, was originally
developed for a normal “beach to offshore” trend. The main parameters of the ichnofacies concept include substrate consistency,
hydrodynamic energy, food supply, food type, salinity, temperature, oxygen levels, sedimentation rate, etc., which tend to change
progressively with increasing water depth. However, the relation to water depth is passive, inasmuch as ichnofacies distribution
is rarely directly controlled only by bathymetry (Frey et al., 1990). Consequently, ichnofacies commonly occur outside the zone
specified in the original paradigm. Continental-shelf ichnofacies are no exception to this observation. (Redrawn from Pemberton
et al., 2001.)

Numerous deposits have been interpreted as tidal shelf sand Ito and Horikawa (2000) interpreted Middle Pleistocene deposits
ridges (e.g., Houthuys and Gullentops, 1988), although the dis- of the Ichijiku Formation of the Boso Peninsula of Japan as a shelf
tinction between estuarine or deltaic tidal sands (see Boyd et al., sand complex formed by the paleo–Kuroshio current. These
this volume) and tidal shelf sand ridges is sometimes difficult deposits reach thicknesses of up to 400 m and are characterized by
(Johnson and Baldwin, 1996). Suter and Clifton (1999) found large-scale trough cross bedding, reactivation surfaces, minor
considerable evidence for the deposition of the Shannon Sand- tabular cross beds, and current ripples. Direct tidal indicators,
stone by tidal currents, including sigmoidal bedding, mud drapes, such as double mud drapes and rhythmites, are lacking.
double mud drapes, elongate rip-up clasts, current ripples, and Recent examples of interpretations of shelf sandbodies in the
rhythmites. These are all features of inshore or reversing tides, subsurface include those of Suter et al. (1996) (Fig. 55); Galloway,
which should not occur prominently in shelf sands deposited by 2002 (Fig. 56); Picarelli et al., (2002); Posamentier, 2002 (Figs. 57–
rotary tidal currents. Quaternary tidal shelf sand ridges show 59); and Handford and Baria, (2003). Galloway (2002), using well
surficial bed forms and internal inclined reflections at various logs and 2D and 3D seismic data, interpreted the Miocene Utsira
scales; thus their interpreted ancient counterparts are character- Formation of the southern North Sea as comprising clastic shelf
ized by extensive cross bedding. Very large-scale cross bedding sands formed in an elongate structural trough, countering alter-
(10–50 m in height), reminiscent of that reported for the Celtic Sea native interpretations that featured submarine fans (Fig. 56).
ridges by Reynaud et al. (1999), has been reported (Surlyk and Storm, tidal, and marine currents combined with abundant sup-
Noe-Nygaard, 1991) from the Cretaceous Raukely Formation of ply of allochthonous sediment off the adjacent platforms to build
East Greenland. Relatively few outcrop examples of ocean-cur- a long-lived aggradational shoal complex. Analogy was made to
rent-deposited sandstones have been reported (Stow et al., 1998). several modern shelf settings, but the Utsira succession is 100–300
376 JOHN R. SUTER

Psilonichnus
Ichnofacies High
Macaronichnus Tide
Assemblage Low
Psilonichnus
Tide
Ichnofacies
Skolithos
Ichnofacies

Suspension Feeding
Fair-weather
Wave Base

Proximal
Cruziana

Macaronichnus
Assemblage Archetypal

Cruziana Ichnofacies
Cruziana

Deposit Feeding
Cruziana
Ichnofacies
SHELF
SHELF

Distal
Grazing & Foraging Cruziana

Storm Wave Base


Zoophycus Ichnofacies

Zoophycus
Ichnofacies

FIG. 45.—An idealized vertical succession for marine softground assemblages from wave-dominated shelf–shoreline deposits, based
on observations from Cretaceous strata from the Western Interior Seaway of North America (redrawn from Pemberton et al., 2001;
modified therein from Pemberton and MacEachern, 1995).

m thick, or several times the thickness of even the largest modern shelf. Utilizing well logs, cores, and beautifully imaged 3D seis-
sand ridges, and it spans about 7 million years. The Utsira mic data, he interpreted a series of linear features as tidal shelf-
succession in this case represents a truly composite system that ridge deposits (Figs. 57–59). These features lie on a transgressive
spans multiple depositional sequences. Although Galloway (2002) ravinement surface (Fig. 59), contain up to 80% net sand, and
argued persuasively for the clastic shelf origin of the Utsira, and range from 0.3 to 2.0 km wide, more than 20 km long, and up to
provided useful summaries of the large-scale architecture and 17 m high. These characteristics are a very good match with those
likely characteristics of clastic shelf shoal systems, the long-lived, of modern ridges, and the comparison gains further support from
basin-centered, repeatedly reoccupied depositional system in- the occurrence of the Miocene features in fields with dozens of
voked has no true modern analog. individual ridges, also a common characteristic on modern shelves.
Posamentier (2002) provided a detailed stratigraphic evalua- Typically, these Miocene ridges are oblique to the reconstructed
tion of Miocene deposits from a portion of the northwest Java paleo-shoreline, contain a sharp edge (possibly the “leading
FACIES MODELS REVISITED: CLASTIC SHELVES 377

A B C

Erosion
D) Erosion and
and Amalagamation
Amalgamation

A)Tempestite
Tempestitedeposition
deposition B)Colonization
Colonizationpost-storm
post-storm C)Fair weather
Fair-weather

“Lam-scram”
FIG. 46.—A) An initial storm event deposits a tempestite, sometimes marked by escape structures of organisms either entrained in
the storm flow or buried by the newly deposited bed. B) Following storm abatement, organisms colonize the sandy substrate,
usually dominated by members of Cruziana or Skolithos ichnofacies. C) With resumed fair-weather deposition, the normal shelf
Zoophycus ichnofacies is reestablished. Subsequent storm events may either substantially erode the fair-weather sediments and
preceding tempestite (D), or preserve enough to produce E) laminated-scrambled or “lam-scram” (Howard, 1975) bedding
(redrawn from Pemberton et al., 2001).

Multiple events; inverse-graded,


bioturbated sand bed

Upward fining storm beds


Low-Concentration
Deposits

High-Concentration
Deposits Poorly sorted clay rich storm
beds;
(River Floods) Flood debris

Multiple events;inverse-graded
bioturbated sand bed
Low-Concentration
Deposits
Upward-fining storm beds

Grain size

FIG. 47.—Schematic drawing illustrating bed architecture on the northern California shelf. Thick, poorly sorted beds deposited by
fluid-mud transport during times of river flooding (high-concentration regimes) alternate with thinner, sandier, bioturbated beds
deposited by wave–current transport in the absence of flood deposits (low-concentration regimes) (redrawn from Fan et al., 2004).
378 JOHN R. SUTER

Laminated siltstone

Climbing, combined-
Combined flow current ripples
Flow FIG. 48.—A general model for a
tempestite deposited by wave-
Asymmetric, modified gravity flows (from
small HCS Myrow et al., 2002).

Massive,
Turbulent graded
Flow

VFS
Silt

MS

CS
Large, well-developed flutes FS

0 1 2 3 30 00’ N 94 00' W
95 00' W
A kilometers 30' Texas
LA

Galveston
Bay

20'

10'
25 m

m
20

38
50'
East Coast, USA
0

Texas
kilometers
m 28 40’ N
0 MAP
10

20' 10' 75 B Map

FIG. 49.—A) Distribution of sand ridges on a portion of the Atlantic continental shelf of the United States and B) the East Texas shelf
in the Gulf of Mexico (redrawn from Snedden and Dalrymple, 1999; after Swift and Field, 1981, Thomas and Anderson, 1994).
FACIES MODELS REVISITED: CLASTIC SHELVES 379

Sea Level
Landward Time 1

Time lines
Nearshore Ridges
Offshore muds
Prograding Barrier Degraded Barrier
Lagoonal muds
Sea Level
Pleistocene substrate Time 2

Barrier and shoreface sands

Aggraded sands Relict Nearshore Ridges

Transgressive Barrier
Degraded
Barriers Sea Level
Time 3

Nearshore Ridges

Relict Nearshore Ridges


FIG. 50.—Schematic diagram for postulated development of New Jersey shelf ridges. Note the existence of regressive and
transgressive shelf ridges as well as degraded barriers (redrawn from Nummedal and Suter, 2002; from Stubblefield et al., 1984).

edge” during their migration), and are asymmetric with the


Flo

thicker parts in the inferred direction of migration (Fig. 58). Core


w

observations reveal that the ridges consist of moderately coarse


Str
ea

sand and contain open-marine fossils and current-formed sedi-


ml

mentary structures, and they are fully encased in muddy sedi-


ine
s

ments immediately above and below (Fig. 59).

A' TOWARDS A CLASTIC-SHELF FACIES MODEL

Considering the complexity of shelf dynamics, the composite


origins of many Holocene and Pleistocene examples, and the
degree of ambiguity associated with outcrop and subsurface
A
interpretations, can we really construct a consistent, commonly
accepted facies model for clastic shelf sands? Johnson and Baldwin
(1996) indicated that it might not be possible to distinguish
between shoreline-associated and offshore sand bodies by facies
analysis alone. Their advice was to concentrate on assemblages of
Ridge data, including texture, sedimentary structures, sandbody geom-
A Crest etry, paleocurrent patterns, and stratigraphic relationships. One
can add that the mineralogy of the sands, the taphonomy of any
Shelf Current
contained shell debris, the biostratigraphy of the encasing and
contained fine-grained sediments, and the ichnology of both the
A A'
shelf sands and muds may be of critical importance in an environ-
mental interpretation. In other words, to make a convincing shelf


B FIG. 51.—Conceptual model for a Shannon Sandstone storm-
dominated shelf sand ridge (redrawn from Swift and Parsons,
1999; modified therein from Gaynor and Swift, 1988), based
Glauconitic Trough Cross-bedded
largely on studies of Atlantic shelf sand ridges. A) Plan view
of lithofacies distribution on a ridge and flow streamlines over
Tabular Cross-bedded Thinly-bedded the ridge crest. B) Flow-parallel cross section through sand-
ridge deposits, showing lithofacies distribution. Migration of
Transverse Bedform such features is oblique to the overall flow direction.
Flow Streamline
380 JOHN R. SUTER

Top of Core

PHOTO X-RAY PHOTO X-RAY PHOTO X-RAY


0.0 m

1.5 m

1.0 m

0.5 m

2.0 m

FIG. 52.—Photographs (left of each pair) and x-radiographs (right of each pair) of vibracores from the landward flank of Peahala Ridge,
offshore New Jersey on the Atlantic continental shelf of North America, illustrating the sedimentary structures and features of
the upper, marine portion of a modern shoreface-attached shelf ridge. Sediments are mostly sand and shell fragments, with grain
size ranging from very fine to coarse. Dominant sedimentary structures are low-angle laminae, with some low-angle cross
bedding apparent around 0.5 m. More cross bedding was found in the actual ridge crest (redrawn from Snedden et al., 1994).

sand interpretation requires an integrated approach involving meaning that the sediment supply to build a shelf sand ridge
multiple datasets. comes from relict deposits, but the ridges themselves are pal-
Given that there are a variety of clastic-shelf facies types, impsest. With increasing submergence, current reworking, and
there can be no single facies model for these deposits. Fine- shelf-ridge migration, the “precursor” is eroded such that once
grained shelf sediments, tempestites, and thin sheet sands over- a ridge has moved a distance greater than its width it has
lying transgressive surfaces are relatively non-controversial evolved to an entirely open-marine depositional system. It is
and are commonly recognized. Recognition of clastic shelf sand immaterial whether the formative currents come from tides,
ridges, volumetrically the most significant type on modern shelves, semipermanent oceanic currents, storm-driven geostrophic
the most controversial deposit in the subsurface and in outcrop, flows, or a combination of any or all of them—only that the
and the most likely hydrocarbon reservoirs, is somewhat more current be sufficiently powerful and of sufficient duration to
problematic. rework a precursor deposit into a shelf sand body. No scale is
Figure 60 shows an evolutionary model for formation of implied in this diagram, but individual ridges on modern shelves
shelf ridges. Following the Huthnance (1982) model, most sand reach thicknesses of up to 40 m.
ridges start life with a “precursor”, a bathymetric perturbation. Shelf ridges can be classified depending upon their progres-
Precursors are very commonly transgressed shoreline features, sion along the evolutionary pathway. Figure 61 schematically
FACIES MODELS REVISITED: CLASTIC SHELVES 381

C
FIG. 53.—Mississippi delta shelf sand bodies. A) Three-stage evolutionary model for the transgression of deltaic complexes in the
Mississippi Delta Plain (redrawn from Penland et al., 1988). Transgression and reworking of deltaic sands, mostly mouth-bar and
beach-ridge-plain deposits, results in progressive formation of wave-dominated spits, barrier islands, and, upon total submer-
gence, shelf sand bodies or shoals. B) Bathymetric map of shelf shoals developed in transgressed, abandoned deltaic complexes
of the Mississippi Delta. C) Detailed bathymetric map of Ship Shoal. D) Transgressive submergence model. This process describes
the progressive reworking of deltaic facies during rising relative base level, illustrating the formation of inner-shelf shoals from
preexisting deltaic shoreline deposits. Vibracores from Ship and Trinity shoals contained substantially different lithofacies,
reflective of their different stages of reworking. Trinity Shoal is currently being buried by the renewed progradation of the
Atchafalaya complex, and will probably be preserved as a composite of barrier-island and shelf sand bodies. Ultimate
preservation of a shelf sand body derived from a preexisting shoreline deposit is thus seen to be a function of the rate of subsidence,
the rate of shoreface translation, marine reworking, and burial.

places the evolutionary spectrum of these ridges above a ravine- alone would not be sufficient to support such a clastic-shelf
ment surface on a continental shelf. Figures 62–64 show concep- interpretation but would be suggestive or supportive. Finer-
tual vertical successions through the different stages of ridge grained deposits, interpreted as deeper-water shelf sediments,
development. A “juvenile” or partially evolved ridge (Types I will overlie the unit.
and II; Figs. 60, 61, 63) would contain characteristics of both its In core or outcrop, an erosional base that may be marked by
precursor and whatever degree of marine reworking occurred a Glossifungites ichnofacies (Figure 54; MacEachern et al., 1992)
before the process clock was turned off. Fully evolved, fully will be overlain by clean, cross-bedded to low-angle laminated
marine, Type III ridges (Fig. 64) contain only deposits laid down sands, depending on the grain size of the ridge. Differing amounts
by shelf processes. of shell material, indicative of depth of formation and reworked
Recognition of such deposits must consider the scale of into the ridge, are present in Quaternary ridges, but the preserva-
observation and data being interpreted. In well logs, ridges tion of carbonate shells through burial diagenesis is problematic.
overlie a sharp basal surface and show an aggradational pat- Authigenic minerals such as glauconite may be present. In storm-
tern, either blocky, upward cleaning, or shaling upward, de- dominated settings, there may be graded storm beds, and possi-
pending on the ridge history (Figs. 56, 57). The log patterns bly hummocky cross stratification. Bed sets should be gently
382 JOHN R. SUTER

0.0 m
Ophiomorpha

Ravinement
Surface

Rangia sp.
Shoal Sand

B
Glossifungites
Ichnofacies
Deltaic Sediments

A
C
1.0 m

FIG. 54.—Features in the Ship Shoal shelf sand body, offshore of the Mississippi Delta (see Figures 10 and 53 for location). A) Portion
of a vibracore through the base of Ship Shoal. A compositie ravinement surface with a Glossifungites ichnofacies separates a 4 m+
thick shelf sand body from muddy sediments of the abandoned Maringouin complex of the Mississippi Delta (see Penland et al.,
1986). In this core the sands lost cohesion during vibracoring, flowed, and became massive. B) Large Ophiomorpha are commonly
found within the upper portions of the sand body, along with macrofauna such as Rangia sp. (Penland et al., 1986). C) Close-up
of the Glosssifungites ichnofacies found beneath the shoal sand (Penland et al., 1986).
FACIES MODELS REVISITED: CLASTIC SHELVES 383

FIG. 55.—Conceptual diagram de-


picting a model for migration
and trapping of gas within the
shoal deposits reworked from
underlying deltaic and/or in-
cised-valley sands. Shelf sands
within the transgressive systems
tract and highstand parasequen-
ces contain most of the gas accu-
mulations in stratigraphic and
combination traps, because of
better top seal relationships (re-
drawn from Suter et al., 1996).
TST/HST Deltaic Deposits This is a well-known play con-
Distributary channel within TST cept for Tertiary deltaic depos-
its in the Gulf of Mexico, some-
Sandy incised-valley fill, LST/TST times colloquially referred to as
the “destructional bar” play. But
Shoal Sand within HST
a “rose by any other name…”
Gas cap the target is often a shelf sand
Gas migration pathways body (see Figure 50).

West East
GR GR Res GR GR Res GR Res GR GR GR

- 500 - 500 - 500 - 500 - 500 - 500 - 500

Plio- Pleistocene

Utsira Formation
- 1000 1000 - 1000 - 1000 - 1000 - 1000 - 1000 - 1000

Oligocene
Lower Miocene

0 20 km
- 1500 - 1500 - 1500 - 1500 - 1500

FIG. 56.—Well-log cross section from the Shetland Platform towards the Horda Platform in the North Sea, showing the log patterns
and regional stratigraphy of the Miocene Utsira Formation. Galloway (2002) interpreted the Utsira as a long-lived, multi-
sequence, shelf shoal system in an elongate epeiric basin. Note the generally blocky, aggradational log stacking pattern above a
basal erosional surface and sharp top boundary, overlain by fine-grained deposits. The formation shows an overall mounded to
downlapping character on seismic data, and numerous features on 3D seismic were interpreted as having formed by unconfined
shelf flows (redrawn from Galloway, 2002).
384 JOHN R. SUTER

A
Northwest Southeast
E-3 E-12 EC-3 EC-1 EE-1 EE-3 EH-2 EH-1 EH-3 E-6

0 Shelf Ridge
Leading Leading
Distributary Channel
Edge Edge
30 m

B
Northwest Southeast
E-3 E-12 EC-1 EE-3 EH-1
EC-3 EE-1 EH-2 EH-3 E-6
448 408 363 351 384 425 439 474 508 558 607 650 661
210 237 258 290 310 318 357 393 422 429 436 454 469
720 720
740 740
760 760
780 780
800 800
820 820
840 840
860 860
880 880
900 900
920 920
940 940
960 960
980 980
100 100
0 0
102 102
0 0
104 104
0 0
106 106
0 0

Shelf Ridge Distributary Channel


100 ms
Leading
Edge
0 1 2 3 km

0 1 2 miles
FIG. 57.—A) Well-log cross section and B) parallel seismic profile through an interpreted shelf ridge in the middle–upper Miocene
Upper Cibulakan Formation from the Northwest Java shelf (redrawn from Posamentier, 2002). Logs show the sand as sharp
based, with an aggradational, somewhat coarsening-upward succession. Seismic profile shows a low-relief feature defined by
amplitude brightening, dimming abruptly at the western (leading) edge, and gradationally at the eastern (trailing) edge.
FACIES MODELS REVISITED: CLASTIC SHELVES 385

Leading
Edge
Leading 5 ms
Edge

10

15

18 ms

0 1.0 km 0 1.0 km
A Seismic Horizon B Isochron Map
Amplitude Extraction
FIG. 58.—Seismic expressions of interpreted shelf ridges from the middle–upper Miocene Upper Cibulakan Formation from the
Northwest Java shelf. A) Seismic amplitude extraction from the reflection corresponding to the shelf ridge. B) Isochron map
between the zero crossings at the top and base of the shelf-ridge seismic reflection. Note that the isochron thick (E) lies along the
sharp western edge of the ridge, presumably its leading edge during migration. (Redrawn from Posamentier, 2002.)

inclined, typically 1–2 degrees, in the direction of ridge migra- of the primary texture and structures of estuarine and marine
tion. Most Quaternary ridges show superimposed bedforms of sediments by tiny, sometimes microscopic organisms living
various scales, so cross bed sets ranging from decimeters to within it, may cause the laminae to be “fuzzy.” Macrofauna and
several meters (and larger) can be expected. Accretionary pack- microfauna contained within the ridge sediments may be reflec-
ages seen in modern shelf ridges (e.g., Fig. 35) will be reflected in tive of the water depth at which they were deposited, unless
internal scour surfaces and changes in bedding orientation. The they have been reworked into the ridge sand. Biostratigraphy of
overall package may be relatively blocky, coarsen upward slightly, the overlying deeper-water shelf deposits will help with the
or even fine upward, depending on the evolution of the ridge and overall context but may not be of significant use in determining
its preservation. the origin of the sand.
Bioturbation may range from rare in continuously high- Fully developed Quaternary shelf sand ridges reach dimen-
energy settings to common or even pervasive in moribund sions that are resolvable on 3D seismic sections, horizon slices,
ridges or in those that were intermittently active. Coupled with and attribute extractions (cf. Posamentier, 2002; Galloway, 2002).
biostratigraphic data, ichnology can be very effective at differ- On horizontal displays, ridges show as long linear features with
entiating fine-grained deposits from shelf, brackish, and fresh- steep leading edges in the direction of ridge migration, while
water settings, but the environments of sand deposition are trailing edges will be less clearly defined (Figs. 57, 58). Ridge
again more problematic (MacEachern et al., 1999). The ichnofa- fields may show as groups of such features, separated by ero-
cies of Quaternary shelf sands have not been systematically sional troughs of various scales. Depending on resolution, shelf
studied, but traces routinely occur in vibracores (e.g., Fig. 54). ridges on seismic-section displays may be mounded, down-
Many of these traces comprise dwelling structures more typical lapping, and composed of internal packages of low-angle clino-
of the Skolithos ichnofacies, or shoreface–foreshore water depths form reflections.
(Fig. 45). This reflects the sandy lithology, relatively high en-
ergy, and well oxygenated conditions preferred by the types of CLASTIC SHELF DEPOSITS IN
organisms that create the burrows comprising the Skolithos SEQUENCE STRATIGRAPHY
ichnofacies. Increased bioturbation at the top of the ridge may
cause an overall fining-upward signature. Cryptic burrowing Because clastic shelves are dependent upon relatively high
or cryptobioturbation (Pemberton et al., 2001), or modification base level for their very existence, it stands to reason that coarse-
386 JOHN R. SUTER

GR
Ravinement
Surface Shelf Sand

Distributary
Channel

ft m
0 0

30 10
FIG. 59.—Core with well log from an interpreted shelf ridge from the middle–upper Miocene Upper Cibulakan Formation from
the northwest Java shelf. In this deposit, from a different ridge than shown in Figures 59 and 60, the log signature is
characterized by a sharp base and fining-upward trend. A Glossifungites ichnofacies marks the contact between upper fine-
grained to medium-grained sandstone above and mudstone below, interpreted as a transgressive ravinement surface
(redrawn from Posamentier, 2002).

grained clastic shelf deposits are more prevalent during rising from sequence stratigraphers?). Controversy does exist over the
and high base-level phases, which are most prevalent within the formation, preservation, and distribution of shelf source rocks
late lowstand, transgressive, and early highstand systems tracts in the rock record, particularly as it relates to ocean anoxia or
(Fig.. 9). Although not commonly invoked as reservoir facies, times of oxygen depletion. However, there is little or no contro-
clastic shelf deposits in general are already in common usage in versy over the fact that they exist and can be recognized, from
sequence stratigraphy. The basal portions of most shoreline facies analysis, biostratigraphy, and geochemistry. A successful
parasequences (sensu Van Wagoner et al., 1990) comprise shelf shelf source facies requires substantial input of organic carbon
deposits (cf. Figs. 62–64). Only through recognition of paleo- material, from either or both terrestrial or in situ sources, and
bathymetric changes as expressed in facies, lithologies, biostra- preservation of the organic carbon from oxidation or
tigraphy, ichnology, and vertical trends can we identify flooding resuspension (Fig. 25). Reduced allochthonous clastic sedimen-
surfaces or parasequence boundaries and construct a sequence tation allows the accumulation of organic material with less
stratigraphic architecture. dilution, resulting in higher concentrations of organic carbon. In
Fine-grained shelf sediments comprise seal and source fa- shelf settings, this is ordinarily best accomplished in the “con-
cies, which generally receive less attention in sequence strati- densed section” (e.g., Loutit et al., 1988), a relatively thin inter-
graphic literature than reservoir deposits. Yet in the shallow val representing a time of reduced deposition, developed dur-
marine realm, these are the most common, volumetrically most ing the period of maximum flooding between the transgressive
significant, and least equivocal manifestation of clastic shelf and highstand systems tracts (Figs. 62–64). In practice, con-
deposits (perhaps because they have received less attention densed sections can occur in association with all flooding sur-
FACIES MODELS REVISITED: CLASTIC SHELVES 387

200 0
Ev
m 0
m Pr
olu
juvenile ridge 5 ec tio
combined flow ur na
so
new marine
sand body r ry FIG. 60.—An evolutionary progres-
P ro sion of shelf sand ridges (redrawn
swale Precursor gr from Snedden and Dalrymple,
es 1999). Following the Huthnance
inlet or incised ravinement surface sio (1982) model, most sand ridges
valley-fill
200 0
n start life with a “precursor”, a
0
Type I m
m bathymetric perturbation that is
combined flow partially evolved 5 most commonly a transgressed
ridge Marine shoreline-associated deposit. Ju-
new marine Sand venile or Type I ridges retain

ac
sand body

cr
much of the characteristics of their

et
swale erosion

io
precursor. With increasing sub-

n
mergence, current reworking,
inlet or incised and shelf ridge migration, the
ravinement surface 400 0
valley-fill 0
fully evolved
m
m “precursor” is eroded and re-
Type II combined flow ridge 5 worked to varying degrees (Type
II, partially evolved). Ultimately,

ac
cr
it may evolve into an entirely open

et
io
erosion marine deposit (Type III or fully-

n
evolved)
ravinement surface

Type III

Increasing terrestrial signature, decreasing seal, source quality

Fig. 62 Fig. 63 Fig. 64


⇓ ⇓ ⇓

Increasing marine signature; increasing source, seal quality

FIG. 61.—A summary diagram for evolution and preservation of clastic shelf sands (redrawn from Penland et al., 1988; Snedden and
Dalrymple, 1999). This highly schematic diagram depicts a set of sand bodies which have developed during a transgression on
an accommodation-dominated shelf, buried by progradation during a subsequent depositional sequence. Arrows and associated
descriptions refer to characteristics of the overlying highstand deposits. Approximate positions of schematic vertical successions
in Figures 62–64 are shown.
388 JOHN R. SUTER

Vertical Depositional
Significant Surfaces Succession Systems Tract
Environment
Flooding Surface Upper Shoreface/
Mouth Bar

Flooding Surface
Highstand

Increasing Terrestrial Palynoflora


Decreasing Marine Microfossils
Flooding Surface Decreasing Source Potential
Decreasing Seal Effectiveness

Flooding Surface

Maximum Flooding Surface Shelf


Sequence Boundary
Shoreface
Highstand

Shelf
SYMBOLS USED IN FIGURES 62-64
Trough cross-beds Sigmoidal beds Bioturbation
Shells
Tabular cross-beds Combined-flow ripples
Low angle laminae Wave ripples Carbonaceous detritus
Plane beds Hummocky cross- Mudclasts
stratification Pebbles
Current ripples

FIG. 62.—A generalized vertical succession through the updip portion of Figure 61. Renewed progradation following the overall
transgression in which the shelf ridges were formed results in the deposition of a stacked highstand systems tract. This could be
part of the same depositional sequence but is herein depicted as separate sequences, with the sequence boundary and maximum-
flooding surface superimposed. Shelf deposits form the basal portions of parasequences, and their recognition is critical to the
identification of flooding surfaces and systems tracts. Recognition of the shelf sediments would presumably be initially made by
abrupt change in lithology and confirmed by biostratigraphic, ichnological, and mineralogical analyses. In this highly schematic
diagram, shelf sediments are progressively less significant volumetrically upward through the succession of the highstand
systems tract. No vertical scale is intended.

faces, but those most likely to encompass significant source Shales and mudrocks are sometimes reservoirs, if the deposit
facies are associated with maximum-flooding surfaces. in question has sufficient organic content and thermal maturity
Clastic-shelf shales can make excellent seal facies. To be a (e.g., Slatt, 2003). Assuming appropriate traps, these rocks can be
successful top, side, or bottom seal, a deposit requires sufficient their own source, seal, and reservoir, with flow typically coming
seal capacity, integrity, and continuity (Fig. 65). Seal capacity is from fractures. Organic-rich shales, such as those which can
the hydrocarbon column height a particular seal bed can sup- occur on clastic shelves, are excellent candidates as fine-grained
port, which is largely a function of its grain size, clay mineral- hydrocarbon systems.
ogy, and porosity. Integrity is expressed as ductility and thick- Once again, clastic-shelf sand reservoirs are more problem-
ness, and largely relates to the ability of shales to flow rather atic. Following their occurrence on Quaternary shelves and given
than suffer brittle fracturing. This is also important in faulting, our current understanding of their formation, clastic-shelf sand
in which shales of insufficient thickness could be offset to a more ridges are most prevalent in the late lowstand, transgressive, and
permeable unit. Continuity is the familiar areal extent, facies early highstand systems tracts of accommodation-dominated
architecture, or distribution of the fine-grained sediment itself. basins. Except for those that comprise erosional remnants, all
Typically the greatest areal extent, continuity, highest clay Quaternary shelf sand ridges overlie a transgressive erosional
content, and lowest coarse-grained lithologic content are found surface (e.g., Fig. 54). This basal surface may be a “simple”
in deposits of the transgressive and early highstand systems ravinement surface or, more likely, a composite surface (cf. Figs.
tracts. Almon et al. (2003) found that shales from the transgres- 61–64). This lower surface may be a sequence boundary, a com-
sive systems tract of the Lewis Shale of South Dakota had posite flooding surface–sequence boundary, a simple flooding
significantly greater seal potential than highstand shales. surface, or a chronostratigraphically insignificant scour surface
FACIES MODELS REVISITED: CLASTIC SHELVES 389

Vertical Depositional
Significant Surfaces Succession Systems Tract
Environment
Inner Shelf
Flooding Surface Highstand

Flooding Surface Increasing Terrestrial Palynoflora


Decreasing Marine Microfossils
Decreasing Source Potential
Decreasing Seal Effectiveness
Flooding Surface
Highstand
Maximum Flooding Surface

Shelf Sand Body


Flooding Surface
Shoreface

Channel Base Diastem Tidal-Inlet Fill


Transgressive

Incised-Valley Fill

Sequence Boundary
Lowstand
Glossifungites Ichnofacies

Shelf
Highstand

FIG. 63.—A hypothetical vertical succession through a nearshore, “juvenile” or partially reworked “Type I” shelf sand ridge (Fig. 61).
This diagram assumes that the deposit has been buried by fine-grained highstand deposition, preserving its character as a shelf
sand body. Such shelf ridges maintain a considerable portion of their precursor deposit, in this case a remnant barrier island
succession with a tidal-inlet fill cutting into an incised valley. No vertical scale is intended.

Vertical Depositional
Significant Surfaces Succession Systems Tract
Environment
Flooding Surface Inner Shelf/
Lower Shoreface Highstand

Flooding Surface

Increasing Terrestrial Palynoflora


Decreasing Marine Microfossils
Flooding Surface Decreasing Source Potential
Decreasing Seal Effectiveness

Flooding Surface

Maximum Flooding Surface Outer Shelf

Shelf Sand Body


Transgressive
Sequence Boundary Glossifungites Ichnofacies

Lower Shoreface Highstand

Shelf

FIG. 64.—A hypothetical vertical succession through a fully marine, fully reworked, “Type III” shelf sand ridge (Fig. 61). This diagram
assumes that the deposit has been buried by fine-grained highstand deposition, preserving its character as a shelf sand body. The
lower contact of the ridge is on a combined flooding surface–sequence boundary. Ridges on Quaternary shelves are characterized
by varying degrees of storm-graded beds, current-formed low-angle laminae, tabular and trough cross bedding, and bioturba-
tion. No vertical scale is intended, but individual Quaternary ridges can reach vertical dimensions of up to 40 m.
390 JOHN R. SUTER

Seal Capacity
Grain size (mineralogy)
Porosity (burial depth)

Successful
Seal

Seal Integrity Seal Continuity


Ductility Facies architecture
Compressibility Thickness
Areal extent

FIG. 65.—A conceptual diagram illustrating the necessary components of bed seal facies quality. If any of the three components
are lacking, a seal will leak or fail. Continental-shelf muds and silty muds commonly have sufficient capacity, integrity, and
continuity to make exceptional top seals (redrawn from Suter and D’Onfro, 2000; after Kaldi and Atkinson, 1997).

(diastem) formed in response to shelf processes and shoal migra- deposits. However, there are several major difficulties in the
tion. The Huthnance–Hulscher model requires a sand supply, application of these facies models to the ancient record: (1) The
currents capable of moving that sand, and a preexisting bottom relatively unusual nature of the late Quaternary—i.e., a time of
irregularity (Snedden and Dalrymple, 1999). Currents of sufficient high-frequency, high-magnitude base-level fluctuations (not to
strength to transport sand are demonstrably available on shelves, mention human influence), (2) the relatively short-lived nature of
and relative sea-level changes provide excellent sources of coarse- the base level cycles, and (3) preservation potential. At various
grained sediments and bottom perturbations in the form of trans- times in the ancient record, however, clastic shelves have been
gressed shorelines. Supply-dominated shelves provide abundant much more prevalent than they are today. It strains credulity to
sediments, mostly mud-dominated. This results in relatively smooth imagine that clastic shelf sands did not develop during extended
bottom conditions, limiting the bottom irregularities necessary for periods of shelf existence, even if we are currently unwilling to
ridge initiation (Swift and Thorne, 1991). However, under the interpret them as such.
appropriate circumstances, such as transgression and reworking Preservation is the most difficult question. We can fully
of an abandoned delta lobe (Penland et al., 1986; Penland et al., characterize the nature of Quaternary shelf deposits—bathym-
1988; Penland et al., 1989), large shelf sand ridges can develop even etry, morphology, lithology, internal architecture, ichnology,
in supply-dominated settings. biostratigraphy—but no matter how thorough our character-
Indeed, given sufficient allochthonous sediment supply and ization, few Quaternary shelf sediments are in their final pro-
sufficiently strong currents, clastic shelf sands need not be re- cess environment. Well studied ridges such as those on the East
stricted to the particular systems tracts. In these circumstances, to Coast of the United States and the North Sea are not likely to
form shelf sands we simply need the shelf to exist; thus they will survive the next eustatic fall in their present forms. During
still be favored in times of relatively high base level. This is well falling base level, shelf deposits will be exposed to renewed
illustrated by the example of Eastern Australia (Boyd et al., 2004a, shoreface processes as well as potential removal by valley
2004b), where, at today’s glacioeustatic highstand, tidal currents incision or other subaerial erosion. Only burial can turn off the
transport sands brought in by longshore drift across the shelf and process clock. This is a conundrum, inasmuch as shelf sands are
onto the continental slope and deeper water (Figs. 38–41). Admit- best developed on accommodation-dominated shelves but their
tedly this is an unusual occurrence, but are we to suppose that no preservation requires burial, i.e., a supply-dominated setting.
other such situations existed in the ancient record? For some insight we can turn to those areas where shelf ridges
are currently being buried, such as the Mississippi Delta in the
FINAL THOUGHTS Gulf of Mexico (Figs. 19, 53). However, these are rapidly evolv-
ing features, even for the Quaternary, and may be uncertain
Quaternary shelf deposits provide excellent models for the analogs for the longer-lived systems which are preserved in the
characteristics of clastic shelf facies, for reservoir, source, and seal stratigraphic record.
FACIES MODELS REVISITED: CLASTIC SHELVES 391

It may be somewhat disquieting to be left with the idea that D.G., Reinson, G.E., Zaitlin, B.A., and Rahmani, R.A., eds., Clastic
clastic shelf sands are complex and their identification is not Tidal Sedimentology: Canadian Society of Petroleum Geologists,
straightforward. Barwis (1989) posed a series of questions rela- Memoir 16, p. 29–39.
tive to our ability to predict occurrences of shelf sands in the ALMON, W.R., DAWSON, W.C., SUTTON, S.J., ETHRIDGE, F.G., AND CASTELBLANCO,
subsurface. Many of those questions still cannot be answered B., 2003, Sequence stratigraphy, facies variation and petrophysical
confidently, but we can draw some insights and perhaps com- properties in deepwater shales, Upper Cretaceous Lewis Shale, south-
fort from the techniques of probabilistic risking and reservoir central Wyoming, in Scott, E.D., Bouma, A.H., and Bryant, W.R., eds.,
modeling employed in the petroleum industry. Instead of ask- Siltstones, Mudstones and Shales: Depositional Processes and Char-
ing: Is this deposit a shelf sand ridge? We can ask: Does the acteristics, SEPM/Gulf Coast Association of Geological Societies,
deposit in question fit what we know of clastic shelf sands? If so, Joint Publication, p. 49–64.
what are the implications of that interpretation for reservoir AMOROSI, A., 1995, Glaucony and sequence stratigraphy: a conceptual
architecture and properties, basin configuration, framework of distribution in siliciclastic sequences: Journal of Sedi-
paleogeographies? It is best to honor the practice of multiple mentary Research v. B65, p. 419–425.
working hypotheses. ARNOTT, R.W., AND SOUTHARD, J.B., 1990, Exploratory flow-duct experi-
ments on combined-flow bed configurations, and some implications
ACKNOWLEDGMENTS for interpreting storm-event stratification: Journal of Sedimentary
Petrology, v. 60, p. 211–219.
I gratefully thank ConocoPhillips for permission to publish ASLAN, A., WHITE, W.A., WARNE, A.G., AND GUEVARA, E., 2003, Holocene
this paper, as well as Julia B. Ericsson, Manager of Sedimentary evolution of the Orinoco Delta, Venezuela: Geological Society of
Systems, for the time and support required for its generation. America, Bulletin, v. 115, p. 479–498.
Henry W. Posamentier and Roger G. Walker solicited this ATKINSON, C.D., GOESTON, M.J.B.G., SPEKSNIJDER, A., AND VAN DER VLUGT, W.,
contribution, and showed remarkable patience and endurance 1986, Storm-generated sandstone in the Miocene Miri Formation,
throughout its intermittent and well-nigh interminable genera- Seria Field, Brunei (N.W. Borneo), in Knight, R.J., and McLean, J.R.,
tion. Once again, the process of producing a review paper has eds., Shelf Sands and Sandstones: Canadian Society of Petroleum
proven truly onerous, but hopefully, ultimately worthwhile. I Geologists, Memoir 11, p. 213–240.
certainly learned a lot. BARWIS, J.H., 1989, The explorationist and shelf sand models: where do we
Over the years I have benefited greatly from discussions and go from here? Gulf Coast Section SEPM Foundation, Seventh Annual
interactions with individuals too numerous to acknowledge Research Conference, Proceedings, p. 1–14.
properly. I would like to particularly thank Henry L. Berryhill, BELDERSON, R.H., 1986, Offshore tidal and non-tidal sand ridges and
Jr., Dag Nummedal, Ron Boyd, and H. Edward Clifton for the sheets: differences in morphology and hydrodynamic setting, in
many hours of discussion and keen insights that they have Knight, R.J., and McLean, J.R,. eds., Shelf Sands and Sandstones:
provided. Canadian Society of Petroleum Geologists, Memoir 11, p. 293–301.
The illustrations in this paper have mostly been drawn BELDERSON, R.H., JOHNSON, M.A., AND KENYON, N.H., 1982, Bedforms, in
from previous work, but several individuals contributed im- Stride, A.H., ed., Offshore Tidal Sands; Processes and Deposits:
agery or data directly. I would like to particularly thank Khalid London, Chapman & Hall, p. 27–55.
Soofi at ConocoPhillips, who created the beautiful satellite BENTLEY, S.J., 2003, Wave–current dispersal of fine-grained fluvial sedi-
altimetry–bathymetry images, and Ron Boyd of the University ments across continental shelves: the significance of hyperpycnal
of Newcastle, Australia, who provided numerous images and plumes, in Scott, E.D., Bouma, A.H., and Bryant, W.R., eds., Siltstones,
photographs, including the stunning bathymetric imagery Mudstones and Shales: Depositional Processes and Characteristics:
from the Australian shelf. Dag Nummedal, then of the Univer- SEPM/Gulf Coast Association of Geological Societies, Joint Publica-
sity of Wyoming and now at Colorado School of Mines, John tion, p. 35–48.
Snedden of ExxonMobil, Thomas Demchuk of ConocoPhillips, BENTLEY, S.J., AND NITTROUER, C.A., 2003, Emplacement, modification and
and Brian Todd of the Geological Survey of Canada Bedford preservation of event stratigraphy on a flood-dominated continental
Institute of Oceanography also provided diagrams and imag- shelf: Eel Shelf, Northern California: Continental Shelf Research, v.
ery. I greatly appreciate their courtesy and assistance. Debbie 23, p. 1465–1493.
Hall of Manzanita Alliances and Hank Tran of ConocoPhillips BERG, R.R., 1975, Depositional environment of Upper Cretaceous Sussex
provided graphics support. The manuscript was immensely Sandstone, House Creek field, Wyoming: American Association of
improved by the reviews of Octavian Catunaneu, David James, Petroleum Geologists, Bulletin, v. 59, p. 2099–2110.
and Glenn Schmidt. Bob Clarke and John Southard assembled BERGMAN, K.M., 1994, Shannon sandstone in Hartzog Draw–Heldt Draw
and edited the final manuscript. I thank them all for their time fields (Cretaceous, Wyoming, USA) reinterpreted as lowstand
and efforts, but I retain responsibility for all remaining errors shoreface deposits: Journal of Sedimentary Research, v. B64, p. 184–
and shortcomings. 201.
Finally, I thank Carmen Fraticelli for her love and long- BERGMAN, K.M., 1999, Cretaceous Sussex Sandstone in House Creek Field
suffering patience. (Wyoming, USA): Transgressive incised shoreface deposits, in
Bergman, K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies:
REFERENCES Sequence Stratigraphic Analysis and Sedimentologic Interpretation:
SEPM, Special Publication 64, p. 297–320.
AIGNER, T., 1985, Storm depositional systems: dynamic stratigraphy in BERGMAN, K.M., AND SNEDDEN, J.W., 1999, Isolated Shallow Marine Sand
modern and ancient shallow-marine sequences, Berlin, Springer- Bodies: Sequence Stratigraphic Analysis and Sedimentologic Inter-
Verlag, Lecture Notes in Earth Sciences, v. 3, 174 p. pretation: SEPM, Special Publication 64, 362 p.
AIGNER, T., AND REINECK, H.E., 1982, Proximality trends in modern storm BERGMAN, K.M., AND WALKER, R.G., 1999, Campanian Shannon Sandstone:
sands from the Helgoland Bight (North Sea) and their implications for an example of a falling stage systems tract deposit, in Bergman, K.M.,
basin analysis: Senckenbergiana Maritima, v. 14, p. 183–215. and Snedden, J.W., eds., Isolated Marine Sandbodies: Sequence Strati-
ALLEN, G.P., 1991, Sedimentary processes and facies in the Gironde graphic Analysis and Sedimentologic Interpretation: SEPM, Special
estuary: A Recent model for macrotidal estuarine systems, in Smith, Publication 64, p. 85–94.
392 JOHN R. SUTER

BERNÉ, S., LERICOLAIS, G., MARSSET, T., BOURILLET, J.-F., AND DE BATIST, M., CURRAY, J.R., 1964, Transgression and regression, in Miller, R.L., ed.,
1998, Erosional offshore sand ridges and lowstand shorefaces: ex- Papers in Marine Geology: New York, The Macmillan Company, p.
amples from tide- and wave-dominated environments of France: 175–203.
Journal of Sedimentary Research, v. 68, p. 540–555. DALRYMPLE, R.W., 1992, Tidal depositional systems, in Walker, R.G., and
BERNÉ, S., TRENTESAUX, A., STOLK, A., MISSIAEN, T., AND DE BATIST, M., 1994, James, N.P., eds., Facies Models: Response to Sea Level Change:
Architecture and long-term evolution of a tidal sand bank: The Geological Association of Canada, p. 195–218.
Middelkerke Bank (southern North Sea): Marine Geology, v. 121, p. DALRYMPLE, R.W., BAKER, E.K., HARRIS, P.T., AND HUGHES, M.G., 2003,
57–72. Geomorphology and sedimentology of the muddy, tide-dominated
BERNÉ, S., VAGNER, P., GUICHARD, F., LERICOLAIS, G., LIU, Z., TRENTESAUX, A., Fly River delta, Papua New Guinea, 2003, in Sidi, F.H., Nummedal,
YIN, P., AND YI, H.I., 2002, Pleistocene forced regressions and tidal sand D., Imbert, P., Darman, H., and Posamentier, H.W., eds., Tropical
ridges in the East China Sea: Marine Geology, v. 188, p. 293–315. Deltas of Southeast Asia—Sedimentology, Stratigraphy, and Petro-
BOESCH, D.F., AND RABALAIS, N.N., 1991, Effects of hypoxia on continental leum Geology: SEPM, Special Publication 76, p. 147–174.
shelf benthos: comparisons between the New York Bight and the Gulf DALRYMPLE, R.W., AND CUMMINGS, D., 2004, The offshore transport of sand
of Mexico, in Tyson, R.V., and Pearson, T.H., eds., Modern and and mud: implications for the determination of wave base and
Ancient Continental Shelf Anoxia: Geological Society of London, shoreline migration distance, in Hampson, G., Steel, R., Burgess, P.,
Special Publication 58, p. 27–34. and Dalrymple, R., eds., Recent Advances in Shoreline–Shelf Stratig-
BOUMA, A.H., BERRYHILL, H.L., BRENNER, R.L., AND KNEBEL, H.J., 1982, raphy (abstract): SEPM Research Conference, Abstracts, p. 5.
Continental shelf and epicontinental seaways, in Scholle, P.A., and DALRYMPLE, R.W., AND HOOGENDOORN, E.L., 1997, Erosion and deposition
Spearing, D., eds., Sandstone Depositional Environments: American on migrating shoreface-attached ridges, Sable Island, Eastern Canada:
Association of Petroleum Geologists, Memoir 31, p. 281–327. Geoscience Canada, v. 24, p. 25–36.
BOYD, R., RUMING, K., DAVIES, S., PAYENBERG, T., AND LANG, S., 2004, Fraser DALRYMPLE, R.W., LEGRESLEY, E.M., FADER, G.B.J., AND PETRIE, B.D., 1992,
Island and Hervey Bay—a classic modern sedimentary environment, The western Grand Banks of Newfoundland: transgressive Holocene
in Boult, P.J., Johns, D.R., and Lang, S.C., eds., Eastern Australian sedimentation under the combined influence of waves and currents:
Basins Symposium II: Petroleum Exploration Society of Australia, Marine Geology, v. 105, p. 95–118.
Special Publication, p. 511–521. DAM, G., AND SURLYK, F., 1998, Stratigraphy of the Neill Klinter Group: a
BOYD, R., RUMING, K., AND ROBERTS, J.J., 2004, Geomorphology and surficial Lower–lower Middle Jurassic tidal embayment succession; Jameson
sediments of the southeast Australian continental margin: Australian Land; East Greenland: Geology of Greenland Survey, Bulletin 175,
Journal of Earth Sciences, v. 51, p. 743–764. 80 p.
BOYLES, J.M., AND SCOTT, A.J., 1982, A model for migrating shelf-bar DAVIS, R.A., JR., AND BALSON, P.S., 1992, Stratigraphy of a North Sea tidal
sandstones in Upper Mancos Shale (Campanian), northwestern Colo- sand ridge: Journal of Sedimentary Petrology, v. 62, p. 116–121.
rado: American Association of Petroleum Geologists, Bulletin, v. 66, DAVIS, R.A., JR., KLAY, J., AND JEWELL, P., 1993, Sedimentology and stratig-
p. 491–508. raphy of tidal sand ridges, Southwest Florida inner shelf: Journal of
BRENNER, R., 1978, Sussex Sandstone of Wyoming—Example of Creta- Sedimentary Petrology, v. 63, p. 91–104.
ceous offshore sedimentation: American Association of Petroleum DE HAAS, H., VAN WEERING, T.C.E., AND DE STIGTER, H., 2002, Organic carbon
Geologists, Bulletin, v. 62, p. 181–200. in shelf seas: sinks or sources, processes and products: Continental
BURNS, B.A., TYE, R.S., KNOCK, D., KREMER, M., GINGRICH, D., SCOTT, A.J., AND Shelf Research, v. 22, p. 691–717.
MACEACHERN, J.A., 2002, The Late Jurassic Alpine C sandstone, a DEMCHUK, T.D., GONZALEZ-GUZMAN, E, GORDON, I.R., CURTIS, C.M., AND
bioturbated, paralic reservoir deposited during a slow transgression: SUTER, J.R., 2004, An integrated chronostratigraphy for the Oficina
North Slope, Alaska (abstract): American Association of Petroleum Formation at Petrozuata, Venezuela: A year 2003 evaluation based on
Geologists–SEPM Annual Meeting, Proceedings. additional stratigraphic and micropaleontological data: Palynology,
CATTANEO, A., CORREGIARI, A., LANGONE, L., AND TRINCARDI, F., 2003, The v. 28, p. 239.
Late-Holocene Gargano subaqueous delta, Adriatic shelf: Sediment DOTT, R.H., JR., AND BOURGEOIS, J., 1982, Hummocky stratification: Sig-
pathways and supply fluctuations: Marine Geology, v. 193, p. 61–91. nificance of its variable bedding sequences: Geological Society of
CHAFETZ, H.S., AND REID, A., 2000, Syndepositional shallow-water precipi- America, Bulletin, v. 32, p. 167–194.
tation of glauconitic minerals: Sedimentary Geology, v. 136, p. 29–42 DRIESE, S.G., FISCHER, M.W., EASTHOUSE, K.A., MARKS, G.T., GOGOLA, A.R.,
CLIFTON, H.E., 1999, Tidal Depositional Systems: ConocoPhillips, internal AND SCHONER, A.E., 1991, Model for genesis of shoreface and shelf
document. sandstone sequences, southern Appalachians: paleoenvironmental
COLE, R.D., AND YOUNG, R.G., 1991, Facies characterization and architec- reconstruction of an Early Silurian shelf system, in Swift, D.J.P.,
ture of a muddy shelf sandstone complex: Mancos B interval of Upper Oertel, G.F., Tillman, R.W., and Thorne, J.A., eds., Shelf Sand and
Cretaceous Mancos Shale, northwest Colorado–northeast Utah, in Sandstone: Geometry, Facies, and Sequence Stratigraphy: Interna-
Miall, A.D., and Tyler, N., eds., The Three-Dimensional Facies Archi- tional Association Sedimentologists, Special Publication 14, p. 309–
tecture of Terrigenous Clastic Sediments and Its Implications for 338.
Hydrocarbon Discovery and Recovery: SEPM, Concepts in Sedimen- DUKE, W.L., 1990, Geostrophic circulation or shallow marine turbidity
tology and Paleontology, v. 3, p. 277–287. currents? The dilemma of paleoflow patterns in storm-influenced
COLE, R.D., YOUNG, R.G., AND WILLIS, G.C., 1997, The Prairie Canyon Mem- prograding shoreline systems: Journal of Sedimentary Petrology, v.
ber, a new unit of the Mancos Shale, west-central Colorado and east- 60, p. 870–883.
central Utah: Utah Geological Survey, Miscellaneous Publication 97-4. DUKE, W.L., ARNOTT, R.W.C., AND CHEEL, R.J., 1991, Shelf sandstones and
COMPTON, J.S., WIGLEY, R., AND MCMILLAN, I.K., 2004, Late Cenozoic hummocky cross stratification: new evidence on a stormy debate:
phosphogenesis on the western shelf of South Africa in the vicinity of Geology, v. 19, p. 625–628.
the Cape Canyon: Marine Geology v. 206, p. 19–40. DYER, K.R., 1995, Sediment transport processes in estuaries, in Perillo,
CURRAY, J.R., 1960, Sediments and history of Holocene transgression, G.M.E., ed., Geomorphology and Sedimentology of Estuaries: Am-
continental shelf, northwest Gulf of Mexico, in Shepard, F.P., Phleger, sterdam, Elsevier, Developments In Sedimentology, v. 53, p. 423–450
F.B., and Van Andel, T.H., eds., Recent Sediments, Northwest Gulf of DYER, K.R., AND HUNTLEY, D.A., 1999, The origin, classification and mod-
Mexico: Tulsa, Oklahoma, American Association of Petroleum Ge- elling of sand banks and ridges: Continental Shelf Research, v. 19, p.
ologists, p. 221–266. 1285–1330.
FACIES MODELS REVISITED: CLASTIC SHELVES 393

EDWARDS, C.M., HODGSON, D.M., FLINT, S.S., AND HOWELL, J.A., 2005, Con- Rocky Mountains: Canadian Journal of Earth Sciences, v. 16, p.
trasting styles of shelf sediment transport and deposition in a ramp 1673–1690.
margin setting related to relative sea-level change and basin floor HAMPSON, G., HOWELL, J.A., AND FLINT, S.S., 1999, A sedimentological and
topography, Turonian (Cretaceous) Western Interior of central Utah, sequence stratigraphic re-interpretation of the Upper Cretaceous
U.S.A: Sedimentary Geology, v. 179, p. 117–152. Prairie Canyon Member (“Mancos B”) and associated strata, Book
EMERY, K.O., 1968, The Atlantic continental shelf and slope of the United Cliffs area, Utah, U.S.A.: Journal of Sedimentary Research, v. 69, p.
States (geologic background): U.S. Geological Survey, Professional 414–433.
Paper 529-A, p. 1–23. HANDFORD, C.R., AND BARIA, L.R., 2003, Exploration potential and high-
EMMEL, F.J., AND CURRAY, J.R., 1982, A submerged late Pleistocene delta and resolution sequence stratigraphy of shelf-sand reservoirs, Miocene,
other features related to sea level changes in the Malacca Strait: South Mississippi, Gulf Coast Association of Geological Societies/
Marine Geology, v. 47, p. 197–216 GCSSEPM, Transactions, v. 53, p. 304–312.
FAAS, R.W., 1991, Rheological boundaries of mud: Where are the limits?: HANSLEY, P.L. AND WHITNEY, C.G., 1990, Petrology, diagenesis, and sedi-
Geo-Marine Letters, v. 11, p. 143–146. mentology of oil reservoirs in Upper Cretaceous Shannon Sandstone
FAN, S., SWIFT D.J.P., TRAYKOVSKI, P., BENTLEY, S., BORGELDD, J.C., REEDE, C.W., Beds, Powder River Basin, Wyoming: U.S. Geological Survey, Bulle-
AND NIEDORODA, A.W., 2004, River flooding, storm resuspension, and tin 1917-C, 33 p.
event stratigraphy on the northern California shelf: observations HARDENBOL, J., THIERRY, J., FARLEY, M.B., JACQUIN, T., DE GRACIANSKY, P.-C.,
compared with simulations: Marine Geology, v. 210, p. 17–41. AND VAIL, P.R., 1998, Mesozoic and Cenozoic sequence chronostrati-
FLEMMING, B.W., 1980, Sand transport and bedform patterns on the conti- graphic framework of European basins, in de Graciansky, P.-C.,
nental shelf between Durban and Port Elizabeth (southeast African Hardenbol, J., Jacquin, T., and Vail, P.R., eds., Mesozoic and Ceno-
continental margin): Sedimentary Geology, v. 26, p. 170–205. zoic Sequence Stratigraphy of European Basins: SEPM, Special
FLEMMING, B.W., 1981, Factors controlling shelf sediment dispersal along Publication 60, p. 3–14, plus charts.
the southeast African continental margin, in Nittrouer, C.A., ed., HARMS, J.C., SOUTHARD, J.B., SPEARING, D., AND WALKER, R.G., 1975, Deposi-
Sedimentary Dynamics of Continental Shelves: Marine Geology, v. tional Environments as Interpreted from Primary Sedimentary Struc-
42, p. 259–277. tures and Stratification Sequences: Society Economic Paleontologists
FRAZIER, D.E., 1974, Depositional episodes: their relationship to the Qua- and Mineralogists, Short Course 2, 161 p.
ternary stratigraphic framework in the northwestern portion of the HARRIS, N.B., 1989, Reservoir geology of Fangst Group (Middle Jurassic),
Gulf basin: Texas Bureau of Economic Geology, Geological Circular Heidrun Field, offshore Mid-Norway: American Association of Pe-
74-1, 28 p. troleum Geologists, Bulletin, v. 73, p. 1415–1435.
FREY, R.W., 1990, Trace fossils and hummocky cross stratification, Upper HARRIS, N.B., ed., 2005, The Deposition of Organic-Carbon-Rich Sedi-
Cretaceous of Utah: Palaios, v. 5, p. 203–218. ments: Models, Mechanisms, and Consequences: SEPM, Special Pub-
GALLOWAY, W.E., 1975, Process framework for describing the morphologic lication 82, 282 p.
and stratigraphic evolution of deltaic depositional systems, in HAYES, M.O., 1967, Hurricanes as Geological Agents: Case Studies of
Broussard, M.L., ed., Deltas, Models for Exploration: Houston, Texas, Hurricanes Carla, 1961, and Cindy, 1963: Texas Bureau of Economic
Houston Geological Society, p. 87–98. Geology, Report of Investigations No. 61, 56 p.
GALLOWAY, W.E., 2002, Paleogeographic setting and depositional architec- HEIN, F., 1987, Tidal/littoral offshore shelf deposits—Lower Cambrian
ture of a sand-dominated shelf depositional system, Miocene Utsira Gog Group, Southern Rocky Mountains, Canada: Sedimentary Geol-
Formation, North Sea Basin: Journal of Sedimentary Research, v. 72, ogy, v. 52, p. 155–182.
p. 476–490. HIGLEY, D.K., PANTEA, M.P., AND SLATT, R.M., 1997, 3-D Reservoir Charac-
GALLOWAY, W.E., AND HOBDAY, D.K., 1996, Terrigenous Clastic Deposi- terization of the House Creek Oil Field, Powder River Basin, Wyo-
tional Systems: Applications to Fossil Fuel and Ground Water Re- ming, V1.00, U.S. Geological Survey, Digital Data Series DDS-33.
sources: Berlin, Springer-Verlag, 488 p. HOOGENDOORN, E.L., AND DALRYMPLE, R.W., 1986, Morphology, lateral
GAYNOR, G.C., AND SCHEIHING, M.H., 1988, Shelf depositional environ- migration and internal structures of shoreface-connected sand ridges,
ments and reservoir characteristics of the Kuparuk River Formation Sable Island Bank, Nova Scotia, Canada: Geology, v. 14, p. 400–403.
(Lower Cretaceous), Kuparuk Field, North Slope, Alaska, in Lomando, HOUBOLT, J.J.H.C., 1968, Recent sediments in the southern Bight of the
A.J., and Harris, P.M., eds., Giant Oil and Gas Fields: SEPM, Core North Sea: Geologie en Mijnbouw, v. 47, p. 245–273.
Workshop no. 12, v. 1, p. 333–389. HOUTHUYS, R., AND GULLENTOPS, F., 1988, The Vlierzele Sands (Eocene,
GAYNOR, G.C., AND SWIFT, D.J.P., 1988, Shannon Sandstone depositional Belgium): a tidal ridge system, in de Boer, P.L., van Gelder, A., and
model: sand-ridge formation on the Campanian Western Interior Nio, S.D., eds., Tide-Influenced Sedimentary Environments and Fa-
Shelf: Journal of Sedimentary Petrology, v. 58, p. 868–880. cies: Dordrecht, The Netherlands, D. Reidel Publishing Co., p. 139–
GOESTEN, M.J.B.G., AND NELSON, P.H., 1992, Draugen Field, in Foster, N.H., 152.
and Beaumont, E.A., eds., Treatise of Petroleum Geology, Structural HOWARD, J. D., 1975, The sedimentological significance of trace fossils, in
Traps: American Association of Petroleum Geologists, vol. 1, p. 37–54. Frey, R.W., ed., The Study of Trace Fossils: New York, Springer-
GOFF, J.A., AUSTIN, J.A., GULICK, S., NORDFJORD, S., CHRISTENSEN, B., Verlag, p. 131–146.
SOMMERFIELD, C., OLSON, H., AND ALEXANDER, C., 2005, Recent and HULSCHER, S.J.M.H., DE SWART, H.E., AND DE VRIEND, H.J., 1993, The genera-
modern marine erosion on the New Jersey outer shelf: Marine Geol- tion of offshore tidal sand banks and sand waves: Continental Shelf
ogy, v. 216, p. 275–296. Research, v. 13, p. 1183–1204.
GOFF, J.A., SWIFT, D.J.P., DUNCAN, C.S., MAYER, L.A., AND HUGHES-CLARKE, J., HUTHNANCE, J. M., 1982, On one mechanism forming linear sand banks:
1999, High resolution swath sonar investigation of sand ridge, dune Estuarine and Coastal Marine Science, v. 14, p. 79–99.
and ribbon morphology in the offshore environment of the New IMBRIE, J., SHACKLETON, N.J., PISIAS, N.G., MORLEY, J.J., PRELL, W.L., MARTINSON,
Jersey margin: Marine Geology, v. 161, p. 307–337. D.G., HAYS, J.D., MACINTYRE, A., AND MIX, A.C., 1984, The orbital theory
GRAY, D.I., 1987, Troll, in Spencer, A.M., Thomas, M.T., Dorc, A.G., Home, of Pleistocene climate: support from a revised chronology of the
P.C., and Larsen, R.M., eds., Geology of the Norwegian Oil and Gas marine δ18O record, in Berger, A., ed., Milankovitch and Climate, Part
Fields: London, Graham & Trotman, p. 389–401. I: Hingham, Massachusetts, Reidel, p. 269–305.
HAMBLIN, A.P., AND WALKER, R.G., 1979, Storm-dominated shallow ma- ITO, M., AND HORIKAWA, K., 2000, Millennial- to decadal-scale fluctuation
rine deposits in the Fernie–Kootenay (Jurassic) transition, southern in the paleo–Kuroshio Current documented in the Middle Pleistocene
394 JOHN R. SUTER

shelf succession on the Boso Peninsula, Japan: Sedimentary Geology, LIU, J.P., MILLIMAN, J.D., GAO, S., AND PENG, C., 2004, Holocene develop-
v. 137, p. 1-8. ment of the Yellow River’s subaqueous delta, North Yellow Sea:
JENNETTE, D.C., AND JONES, C.R., 1995, Sequence stratigraphy of the Upper Marine Geology v. 209, p. 45–67.
Cretaceous Tocito Sandstone: a model for tidally influenced incised LIU, Z.X., DIAS, D.X., BERNÉ, S., YANG, W.K., MARSSET, T., TANG, Y.X., AND.
valleys, San Juan Basin, New Mexico, in Van Wagoner, J.C., and BOURILLET, J.F., 1998, Tidal depositional systems of China’s continental
Bertram, G.T., eds., Sequence Stratigraphy of Foreland Basin Depos- shelf, with special reference to the eastern Bohai Sea: Marine Geology,
its: American Association of Petroleum Geologists, Memoir 64, p. v. 145, p. 225–268.
311–348. LOUTIT, T.S., HARDENBOL, J., AND VAIL, P.R., 1988, Condensed sections: the
JOHNSON, H.D., AND BALDWIN, C.T., 1996, Shallow clastic seas, in Reading, key to age determination, an integrated approach, in Wilgus, C.K.,
H.G., ed., Sedimentary Environments; Processes, Facies and Stratig- Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and
raphy: Oxford, U.K., Blackwell Science, p. 232–280. Van Wagoner, J.C., eds., Sea Level Changes: An Integrated Approach:
JOHNSON, K.S., PAULL, C.K., BARRY, J.P., AND CHAVEZ, F.P., 2001, A decadal SEPM, Special Publication 42, p. 188–213.
record of underflows from a coastal river into the deep sea: Geology, MACEACHERN, J.A., RAYCHAUDHURI, I., AND PEMBERTON, S.G., 1992, Strati-
v. 29, p. 1019–1022. graphic applications of the Glossifungites ichnofacies: delineating
KALDI, J.G., AND ATKINSON, C.D., 1997, Evaluating seal potential: example discontinuities in the rock record, in Pemberton, S.G., ed., Application
from the Talang Akar Formation, offshore northwest Java, Indonesia, of Ichnology to Petroleum exploration: A Core Workshop: SEPM,
in Surdam, R.C., ed., Seals, Traps, and the Petroleum System: Ameri- Core Workshop 17, p. 169–198.
can Association of Petroleum Geologists, Memoir 67, p. 85–101. MACEACHERN, J.A., STELCK, C.R., AND PEMBERTON, S.G., 1999, Marine mud-
KELLER, G.H., AND RICHARDS, J.F., 1967, Sediments of the Malacca Strait, stone deposition: paleoenvironmental interpretations, in Bergman,
Southeast Asia: Journal of Sedimentary Petrology, v. 37, p. 102–127. K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies: Sequence
KHAN, S., IMRAN, J., BRADFORD, S., AND SYVITSKI, J.A., 2005, Numerical model- Stratigraphic Analysis and Sedimentologic Interpretation: SEPM,
ing of hyperpycnal plume: Marine Geology, v. 222–223, p. 193–211. Special Publication 64, p. 205–225.
KINEKE, G.C., STERNBERG, R.W., TROWBRIDGE, J.H., AND GEYER, W.R., 1996, MCBRIDE, R.A., ANDERSON, L.C., TUDORAN, A., ROBERTS, H.H., SEN GUPTA,
Fluid-mud processes on the Amazon continental shelf: Continental B.K., AND BYRNES, M.R., 1999, Holocene stratigraphic architecture of a
Shelf Research, v. 16, p. 667–696. sand-rich shelf and origin of linear shoals: northeastern Gulf of
KINEKE, G.C., WOOLFE, K.J., KUEHL, S.A., MILLIMAN, J.D., DELLAPENNA, T.M., Mexico, in Bergman, K.M., and Snedden, J.W., eds., Isolated Marine
AND PURDON, R.G., 2000, Sediment export from the Sepik River, Papua Sandbodies: Sequence Stratigraphic Analysis and Sedimentologic
New Guinea: evidence for a divergent sediment plume: Continental Interpretation: SEPM, Special Publication 64, p. 95–126.
Shelf Research, v. 20, p. 2239–2266. MCBRIDE, R.A., AND MOSLOW, T.F., 1991, Origin, evolution, and distribution
KITAMURA, A., 1998, Glaucony and carbonate grains as indicators of the of shoreface sand ridges, Atlantic inner shelf, U.S.A.: Marine Geology,
condensed section: Omma Formation, Japan: Sedimentary Geology, v. 97, p. 57–85.
v. 122, p. 151–163. MEIJER-DREES, N.C., AND MYHR, D.W., 1981, The Upper Cretaceous Milk
KRAUSE, F.F., COLLINS, H.N., NELSON, D.A., MACHEMER, S.D., AND FRENCH, River and Lea Park formations in southeastern Alberta: Bulletin of
P.R., 1987, Multiscale anatomy of a reservoir: geological characteriza- Canadian Petroleum Geology, v. 29, p. 42–74.
tion of Pembina–Cardium Pool, west-central Alberta, Canada: Ameri- MILLER, R.G., 1990, A paleoceanographic approach to the Kimmeridge
can Association of Petroleum Geologists, Bulletin, v. 71, p. 1233–1260. Clay Formation, in Huc, A.Y., ed., Deposition of Organic Facies:
KRAUSE, F., AND NELSON, D.A., 1991, Evolution of an Upper Cretaceous American Association of Petroleum Geologists, Studies in Geology
(Turonian) shelf sandstone ridge in the Cardium Formation, Pembina 30, p. 13–26.
area, west-central Alberta, Canada, in Swift, D.J.P., Oertel, G.F., MORRIS, W.R., HANNON, R.C., HELMOLD, K.P., KNOCK, D.G., AND POSAMENTIER,
Tillman, R.W., and Thorne, J.A., eds., Shelf Sand and Sandstone H.W., 2000, Jurassic Alpine sandstone, North Slope, Alaska: an ex-
Bodies; Geometry, Facies and Sequence Stratigraphy: International ample of shoreface deposition within a punctuated transgressive
Association of Sedimentologists, Special Publication 14, p. 427–456. succession (abstract): American Association of Petroleum Geolo-
LE BOT, S., AND TRENTESAUX, A., 2004, Types of internal structure and gists–SEPM, Annual Meeting, Proceedings.
external morphology of submarine dunes under the influence of tide- MORTON, R.A., 1981, Formation of storm deposits by wind-forced currents
and wind-driven processes (Dover Strait, northern France): Marine in the Gulf of Mexico and the North Sea, in Nio, S.D., Shuttenhelm,
Geology, v. 211, p. 143–168. R.T.E., and van Weering, Tj.C.E., eds., Holocene Marine Sedimenta-
LECKIE, D.A., 1988, Wave-formed, coarse-grained ripples and their rela- tion in the North Sea Basin: International Association of Sedimentolo-
tionship to hummocky cross-stratification: Journal of Sedimentary gists, Special Publication 5, p. 385–396.
Petrology, v. 58, p. 607–622. MULDER, T., AND SYVITSKI, J.P.M., 1995, Turbidity currents generated at
LECKIE, D.A., AND KRYSTINIK, L.F., 1989, Is there evidence for geostrophic river mouths during exceptional discharges to the world oceans:
currents preserved in the sedimentary record of inner to middle-shelf Journal of Geology, v. 103, p. 285–299.
deposits?: Journal of Sedimentary Petrology, v. 59, p. 862–870. MULLER, J., 1959, Palynology of Recent Orinoco delta and shelf sediments:
LECKIE, D.A., AND POTOCKI, D.J., 1998, Sedimentology and petrography of Reports of the Orinoco Shelf Expedition; Volume 5: Micropaleontol-
marine shelf sandstones of the Cretaceous, Scatter and Garbutt for- ogy, v. 5, p. 1–32.
mations, Liard Basin, northern Canada: Bulletin of Canadian Petro- MYROW, P.M., FISCHER, W., AND GOODGE, J.W., 2002, Wave-modified turbid-
leum Geology, v. 46, p. 30–50. ites: combined-flow shoreline and shelf deposits, Cambrian, Antarc-
LEEDER, M., 1999, Sedimentology and Sedimentary Basins; From Turbu- tica: Journal of Sedimentary Research, v. 72, p. 641–656.
lence to Tectonics: Oxford, U.K., Blackwell Science, 592 p. MYROW, P.M., AND SOUTHARD, J.B., 1996, Tempestite deposition: Journal of
LEVELL, B.K., 1980, A Late Precambrian tidal shelf deposit, the lower Sedimentary Research, v. 66, p. 875–887.
Sandfjord Formation, Finnmark, North Norway: Sedimentology, v. NEILL, C.F., AND ALLISON, M.A., 2005, Subaqueous deltaic formation on the
27, p. 539–557. Atchafalaya Shelf, Louisiana: Marine Geology, v. 214, p. 411–430.
LI, M.Z., AND AMOS, C.L., 1999, Sheet flow and large wave ripples under NELSON, H.F., AND BRAY, E.B., 1970, Stratigraphy and history of the Ho-
combined waves and currents: field observations, model predictions locene sediments in the Sabine–High Island area, Gulf of Mexico, in
and effects on boundary layer dynamics: Continental Shelf Research, Morgan, J.P., ed., Deltaic Sedimentation, Modern and Ancient: SEPM,
v. 19, p. 637–663. Special Publication 15, p. 48–77.
FACIES MODELS REVISITED: CLASTIC SHELVES 395

NIO, S.D., AND YANG, C., 1991, Diagnostic attributes of clastic tidal depos- forms in the Central Basin, Spitsbergen: Journal of Sedimentary
its: a review, in Smith, D.G., Reinson, G.E., Zaitlin, B.A., and Rahmani, Research, v. 71, p. 897–914.
R.A., eds., Clastic Tidal Sedimentology: Canadian Society of Petro- PLINK-BJÖRKLUND, P., AND STEEL, R.J., 2004, Initiation of turbidity currents:
leum Geologists, Memoir 16, p. 3–28. outcrop evidence for Eocene hyperpycnal flow turbidites: Sedimen-
NITTROUER, C.A., KUEHL, S.A., DEMASTER, D.J., AND KOWSMANN, R.O., 1986, tary Geology, v. 165, p. 29–52.
The deltaic nature of Amazon shelf sedimentation: Geological Society PLINT, A.G., 1988, Sharp-based shoreface sequences and “offshore bars” in
of America, Bulletin, v. 97, p. 444–458. the Cardium Formation of Alberta: their relationship to relative
NITTROUER, C.A., AND STERNBERG, R.W., 1981, The formation of sedimentary changes in sea level, in Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C.,
strata in an allochthonous shelf environment: The Washington conti- Posamentier, H.W., Ross, C.A., and Van Wagoner, J.C., eds., Sea Level
nental shelf: Marine Geology., v. 42, p. 201–232. Changes: An Integrated Approach: SEPM, Special Publication 42, p.
NITTROUER, C.A., AND WRIGHT, L.D., 1994, Transport of particles across 357–370.
continental shelves: Reviews of Geophysics, v. 32, p. 85–113. PLINT, A.G., AND NUMMEDAL, D., 2000, The falling stage systems tract:
NIXON, R.P., 1973, Oil source beds in Cretaceous Mowry Shale of north- recognition and importance in sequence stratigraphic analysis, in
western interior United States: American Association of Petroleum Hunt, D., and Gawthorpe, R.L., eds., Sedimentary Responses to
Geologists, Bulletin, v. 57, p. 136–161. Forced Regressions: Geological Society of London, Special Publica-
NOTTVEDT, A., AND KREISA, R.D., 1987, Model for the combined flow origin tion 172, p. 1–17.
of hummocky cross-stratification: Geology, v. 15, p. 357–361. POPE, D.L., PENLAND, S., SUTER, J.R., AND MCBRIDE, R.A., 1990, Holocene
NUMMEDAL, D., AND RILEY, G.W., 1999, The origin of the Tocito Sandstone geologic framework of the Trinity Shoal region, Louisiana continental
and its sequence stratigraphic lessons, in Bergman, K.M., and Snedden, shelf: Gulf Coast Section SEPM, 12th Annual Research Conference,
J.W., eds., Isolated Shallow Marine Sand Bodies: Sequence Strati- Proceedings, p. 191–201.
graphic Analysis and Sedimentologic Interpretation: SEPM, Special PORTER-SMITH, R., HARRIS, P.T., ANDERSEN, O.B., COLEMAN, R., GREENSLADE,
Publication 64, p. 227–254. D., AND JENKINS, C.J., 2004, Classification of the Australian continental
NUMMEDAL, D., AND SUTER, J.R., 2002, Continental shelf sand ridges: gen- shelf based on predicted sediment threshold exceedance from tidal
esis, stratigraphy, and petroleum significance, in 22nd Annual Gulf currents and swell waves: Marine Geology, v. 211, p. 1–20.
Coast SEPM Foundation Bob F. Perkins Research Conference, p. 503– POSAMENTIER, H.W., 2002, Ancient shelf ridges—A potentially significant
518. component of the transgressive systems tract: Case study from off-
ODIN, G.S., AND MATTER, A., 1981, De glauconarium origine: Sedimentol- shore northwest Java: American Association of Petroleum Geolo-
ogy, v. 28, p. 611–641. gists, Bulletin, v. 86, p. 75–106.
PARSONS, J.D., BUSH, J.W.M., AND SYVITSKI, J.P.M., 2001, Hyperpycnal plume POSAMENTIER, H.W., ALLEN, G.P., JAMES, D.P., AND TESSON, M., 1992, Forced
formation with small sediment concentrations: Sedimentology, v. 48, regressions in a sequence stratigraphic framework: concepts, ex-
p. 465–478. amples, and exploration significance: American Association of Petro-
PATTISON, S.A.J., 2005, Isolated highstand shelf sandstone body of turbid- leum Geologists, Bulletin, v. 76, p. 1687–1709.
itic origin, lower Kenilworth Member, Cretaceous Western Interior, POUMOT, C., 1989, Palynological evidence for eustatic events in the tropical
Book Cliffs, Utah, USA: Sedimentary Geology, v. 177, p. 131–144. Neogene: Centres de Recherches Exploration-Production Elf
PEMBERTON, S.G., AND MACEACHERN, J.A., 1995, The sequence strati- Aquitaine, Bulletin, v. 13, p. 437–453.
graphic significance of trace fossils: examples from the Cretaceous RABALAIS, N. N., TURNER, R.E., AND WISEMAN, W.J., 2001, Hypoxia in the Gulf
foreland basin of Alberta, Canada, in Van Wagoner, J.C., and Bertram, of Mexico: Journal of Environmental Quality, v. 30, p. 320–329.
G., eds., Sequence Stratigraphy of Foreland Basin Deposits—Out- RABALAIS, N.N., TURNER, R.E., WISEMAN, W.J., AND BOESCH, D.F., 1991, A brief
crop and Subsurface Examples from the Cretaceous of North summary of hypoxia on the northern Gulf of Mexico continental shelf:
America: American Association of Petroleum Geologists, Memoir 1985–1988, in Tyson, R.V., and Pearson, T.H., eds., Modern and
64, p. 429–475. Ancient Continental Shelf Anoxia: Geological Society of London,
PEMBERTON, S.G., SPILA, M., PULHAM, A.J., SAUNDERS, T., MACEACHERN, J.A., Special Publication 58, p. 35–47.
AND SINCLAIR, I., 2001, Ichnology and Sedimentology of Shallow to RAMSAY, P.J., 1994, Marine geology of the Sodwana Bay shelf, southeast
Marginal Marine Systems, Ben Nevis and Avalon Reservoirs, Jeanne Africa: Marine Geology, v. 120, p. 225–247.
D’Arc Basin: Geological Association of Canada, Short Course Notes, REY, J., AND HIDALGO, M.C., 2004, Siliciclastic sedimentation and sequence
vol. 15, 343 p. stratigraphic evolution on a storm-dominated shelf: the Lower Or-
PENLAND, S., BOYD, R.L., AND SUTER, J.R., 1988, Transgressive depositional dovician of the Central Iberian Zone (NE Jaén, Spain): Sedimentary
systems of the Mississippi delta plain: a model for barrier shoreline Geology, v. 164, p. 89–104.
and shelf sand development: Journal of Sedimentary Petrology, v. 58, REYNAUD, J.-Y., BERNADETTE, T., PROUST, J-N.L., DALRYMPLE, R.W., MARSSET,
p. 932–949. T., DE BATIST, M., BOURILLET, J-F., AND LERICOLAIS, G., 1999, Eustatic and
PENLAND, S., SUTER, J.R., MCBRIDE, R.A., WILLIAMS, S.J., KINDINGER, J.L., AND hydrodynamic controls on the architecture of a deep shelf sand bank
BOYD, R., 1989, Holocene sand shoals offshore of the Mississippi River (Celtic Sea): Sedimentology, v. 46, p. 703–721
Delta plain: Gulf Coast Association of Geological Societies, Transac- RODRIGUEZ, A.B., AND ANDERSON, J. B., 2004, Contourite origin for shelf and
tions, v. 39, p. 471–480. upper slope sand sheet, offshore Antarctica: Sedimentology v. 51, p.
PENLAND, S., SUTER, J.R., AND MOSLOW, T.F., 1986, Inner-shelf shoal sedimen- 699–711
tary facies and sequences: Ship Shoal, Northern Gulf of Mexico, in RODRIGUEZ, A.B., ANDERSON, J.B., SIRINGAN, F.P., AND TAVIANI, M., 1999,
Moslow, T.F., and Rhodes, E.G., eds., Modern and Ancient Shelf Sedimentary facies and genesis of Holocene sand banks on the east
Clastics: A Core Workshop: SEPM, Core Workshop 9, p. 73–122. Texas inner continental shelf, in Bergman, K.M., and Snedden, J.W.,
PICARELLI, A.T., SAVINI, R.R., ABREU, V., GROSO, S., ARGUELLO, J., AND SALAS, eds., Isolated Shallow Marine Sandbodies: Sequence Stratigraphic
D., 2002, Sequence stratigraphic evolution of the Oficina Formation, Analysis and Sedimentologic Interpretation: SEPM, Special Publica-
eastern Venezuela Basin: Depositional systems and sand body geom- tion 64, p. 165–178.
etry in a non-marine to estuarine environment: American Association ROY, P.S., CROWELL, P.J., FERLAND, M.A., AND THOM, B.G., 1994, Wave-
of Petroleum Geologists, Bulletin, v. 86, Supplement. dominated coasts, in Carter, R.W.G., and Woodroffe, C.D., eds., Coastal
PLINK-BJÖRKLUND, P., MELLERE, D., AND STEEL, R.J., 2001, Turbidite variabil- Evolution: Late Quaternary Shoreline Morphodynamics: Cambridge,
ity and architecture of sand-prone, deepwater slopes: Eocene clino- U.K., Cambridge University Press, p. 121–186.
396 JOHN R. SUTER

RINE, J.M., AND GINSBURG, R.N., 1985, Depositional Environments of a mud A.H., ed., Offshore Tidal Sands: Processes and Deposits: London,
shoreface in Suriname, South America—a mud analogue to sandy, Chapman & Hall, p. 95–125.
shallow marine deposits: Journal of Sedimentary Petrology, v. 55, p. STUBBLEFIELD, W.L., MCGRAIL, D.W., AND KERSEY, D.G., 1984, Recognition of
633–652. transgressive and post-transgressive sand ridges on the New Jersey
RINE, J.M., TILLMAN, R.W., CULVER, S.J., AND SWIFT, D.J.P., 1991, Generation continental shelf, in Tillman, R.W., and Siemers, C.T., eds., Siliciclastic
of Late Holocene ridges on the middle continental shelf of New Jersey, Shelf Sediments: SEPM, Special Publication 34, p. 1–23.
USA—Evidence for formation in a mid-shelf setting based upon SULLIVAN, M.D., VAN WAGONER, J.C., JENNETTE, D.C., FOSTER, M.E., STUART,
comparison with a nearshore ridge, in Swift, D.J.P. , Oertel, G.F., R.M., LOVELL, R.W., AND PEMBERTON, S.G., 1997, High resolution se-
Tillman, R.W., and Thorne, J.A., eds., Shelf Sand and Sandstone quence stratigraphy and architecture of the Shannon Sandstone,
Bodies; Geometry, Facies and Sequence Stratigraphy: International Hartzog Draw Field, Wyoming, in Gulf Coast Section SEPM, 18th
Association Sedimentologists, Special Publication 14, p. 395–426. Annual Research Conference, p. 331–344.
SCHEIHING, M.H., AND GAYNOR, G.C., 1991, The shelf sand-plume model: a SURLYK, F., AND NOE-NYGAARD, N., 1991, Sand bank and dune facies
critique: Sedimentology, v. 38, p. 433–444. architecture of a wide intracratonic seaway: late Jurassic–Early Creta-
SCHWARZKOPF, T.A., 1993, Model for prediction of organic carbon content ceous Raukelv Formation, Jameson Land, East Greenland, in Miall,
in possible source rocks: Marine and Petroleum Geology, v. 10, p. 478– A.D., and Tyler, N., eds., The Three-Dimensional Facies Architecture
492. of Terrigenous Clastic Sediments and Its Implications for Hydrocar-
SEILACHER, A., 1967, Bathymetry of trace fossils: Marine Geology, v. 5, p. bon Discovery and Recovery: SEPM, Concepts in Sedimentology and
413–428. Paleontology, v. 3, p. 261–276.
SMITH, J.D., 1970, Stability of a sand bed subjected to a shear flow of low SUTER, J.R., 1987, Ancient fluvial systems and Holocene deposits, south-
Froude number: Journal of Geophysical Research, v. 75, p. 5428–5940. western Louisiana continental shelf, in Berryhill, H.L., Jr., ed., Late
SNEDDEN, J.W., AND BERGMAN, K.M., 1999, Isolated shallow marine sand Quaternary Facies and Structure, Northern Gulf of Mexico: American
bodies: deposits for all interpretations, in Bergman, K.M., and Snedden, Association of Petroleum Geologists, Studies in Geology 23, p. 81–
J.W., eds., Isolated Shallow Marine Sandbodies: Sequence Strati- 129.
graphic Analysis and Sedimentologic Interpretation: SEPM, Special SUTER, J.R., 2001, Sequence Stratigraphy in a Development Setting:
Publication 64, p. 1–11. Petrozuata, Faja Petrolifera del Orinoco, Venezuela: American Asso-
SNEDDEN, J.W., AND DALRYMPLE, R.W., 1999, Modern shelf sand bodies: ciation of Petroleum Geologists, Distinguished Lecture 2001/2002.
from historical perspective to a unified theory for sand body genesis SUTER, J.R., 2003, Late Quaternary shelf margin deltas, northwest Gulf of
and evolution, in Bergman, K.M., and Snedden, J.W., eds., Isolated Mexico, in Gulf Coast Section SEPM, 34th Annual Bob F. Perkins
Shallow Marine Sandbodies: Sequence Stratigraphic Analysis and Research Conference, p. 200–259.
Sedimentologic Interpretation: SEPM, Special Publication 64, p. 13–28. SUTER, J.R., ABDULAH, K., TRAVIS, P.D., GLAGOLA, P.A., YOUNG, J.W., WATSO,
SNEDDEN, J.W., AND JUMPER, R.S., 1990, Shelf and shoreface reservoirs, Tom D.C., AND FINLEY, W.E, 1996, An integrated seismic and well-log
Walsh–Owen Field, Texas, in Barwis, J.H., McPherson, J.G., and sequence stratigraphic study over 32 Grand Isle/West Delta OCS
Studlick, J.R.J., eds., Sandstone Petroleum Reservoirs: New York, Blocks: Gulf Coast Section SEPM, 17th Annual Research Conference,
Springer-Verlag, Casebooks in Earth Sciences, p. 415–436. p. 305–314.
SNEDDEN, J.W., KREISA, R.D., TILLMAN, R.W., CULVER, S.J., AND SCHWELLER, SUTER, J.R., BERRYHILL, H.L., JR., AND PENLAND, S., 1987, Late Quaternary sea
W.J., 1999, An expanded model for modern shelf sand ridge genesis level fluctuations and depositional sequences, southwest Louisiana
and evolution on the New jersey Atlantic shelf, in Bergman, K.M., and continental shelf, in Nummedal, D., Pilkey, O.H., and Howard. J.D.,
Snedden, J.W., eds., Isolated Marine Sandbodies: Sequence Strati- eds., Sea-Level Changes and Coastal Evolution: SEPM, Special Pub-
graphic Analysis and Sedimentologic Interpretation: SEPM, Special lication 41, p. 199–222.
Publication 64, p. 147–163. SUTER, J.R. AND CLIFTON, H.E., 1999, The Shannon Sandstone and isolated
SNEDDEN, J.W., KREISA, R.D., TILLMAN, R.W., SCHWELLER, W.J., CULVER, S.J., linear sand bodies: interpretations and realizations, in Bergman,
AND WINN, R.D., 1994, Stratigraphy and genesis of a modern shoreface- K.M., and Snedden, J.W., eds., Isolated Shallow Marine Sand Bodies:
attached sand ridge, Peahala Ridge, New Jersey: Journal of Sedimen- Sequence Stratigraphic Analysis and Sedimentologic Interpretation:
tary Research, v. B64, p. 560–581. SEPM, Special Publication 64, p. 321-356.
SNEDDEN, J.W., AND NUMMEDAL, D., 1991, Origin and geometry of storm- SUTER, J.R., AND D’ONFRO, P., 2000, Sequence stratigraphy and seal quality
deposited sand beds in modern sediments of the Texas continental prediction in Block B, Natuna Sea: the relationship between top seal
shelf, in Swift, D.J.P., Tillman, R.W., and Oertel, G.F., eds., Shelf Sand quality, lithofacies, and depositional environment: American Asso-
and Sandstone Bodies; Geometry, Facies and Sequence Stratigraphy: ciation of Petroleum Geologists, Bulletin, v. 84, p. 1395–1518.
International Association Sedimentologists, Special Publication 14, p. SWIFT, D.J.P., 1972, Implications of sediment dispersal from bottom cur-
283–308. rent measurements: some specific problems in understanding bottom
SNEDDEN, J.W., NUMMEDAL, D., AND AMOS, A.F., 1988, Storm- and fair- sediment distribution and dispersal on the continental shelf: a discus-
weather combined flow on the Central Texas continental shelf: Jour- sion of two papers, in Swift, D.J.P., Duane, D.B., and Pilkey, O.H., eds.,
nal of Sedimentary Petrology, v. 58, p. 580–595. Shelf Sediment Transport; Process and Pattern: Stroudsberg, Penn-
SPEARING, D.R., 1976, Upper Cretaceous Shannon Sandstone: an offshore sylvania, Dowden, Hutchinson & Ross, p. 363–371.
shallow-marine sandbody: Wyoming Geological Association, Guide- SWIFT, D.J.P., 1975, Tidal sand ridges and shoal retreat massifs: Marine
book, p. 65–72. Geology, v. 18, p. 105–134.
STONECIPHER, S.A., 1999, Genetic characteristics of glauconite and siderite: SWIFT, D.J.P., AND FIELD, M.F., 1981, Evolution of a classic ridge field,
implications for the origin of ambiguous isolated marine sand bodies, Maryland sector, North American inner shelf: Sedimentology, v. 28,
in Bergman, K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies: p. 461–482.
Sequence Stratigraphic Analysis and Sedimentologic Interpretation: SWIFT, D.J.P., FIGUEIREDO, A.G., FREELAND, G.L., AND OERTEL, G.F., 1983,
SEPM, Special Publication 64, p. 191–204. Hummocky cross-stratification and megaripples: a geological double
STOW, D.A., FAUGERES, J-C., VIANA, A., AND GONTHIER, E., 1998, Fossil standard?: Journal of Sedimentary Petrology, v. 53, p. 1295–1318.
contourites: a critical review: Sedimentary Geology, v. 115, p. 3–31. SWIFT, D.J.P., HAN, G., AND VINCENT, C.E., 1986, Fluid processes and sea-
STRIDE, A.H., BELDERSON, R.H., KENYON, N.H., AND JOHNSON, M.A., 1982, floor response on a modern storm-dominated shelf: Middle Atlantic
Offshore tidal deposits: sand sheet and sand bank facies, in Stride, shelf of North America: Part 1. The storm-current regime, in Knight,
FACIES MODELS REVISITED: CLASTIC SHELVES 397

R.J., and McLean, J.R., eds., Shelf Sands and Sandstones: Canadian VIANNA, M.L., SOLEWICZ, R., CABRAL, A.P., AND TESTA, V., 1991, Sandstream
Society of Petroleum Geologists, Memoir 11, p. 99–119. on the northeast Brazilian shelf: Continental Shelf Research, v. 11, p.
SWIFT, D.J.P., AND PARSONS, B.S., 1999, Shannon Sandstone of the Powder 509–524.
River Basin: Orthodoxy and revisionism in stratigraphic thought, in WALKER, R.G., 1984. Shelf and shallow marine sands, in Walker, R.G., ed.,
Bergman, K.M., and Snedden, J.W., eds., Isolated Marine Sandbodies: Facies Models: Geological Association of Canada, p. 141–170.
Sequence Stratigraphic Analysis and Sedimentologic Interpretation: WALKER, R.G., AND BERGMAN, K.M., 1993, Shannon sandstone in Wyoming:
SEPM, Concepts in Sedimentology and Paleontology, v. 6, p. 55–84. a shelf ridge complex reinterpreted as lowstand shoreface deposits:
SWIFT, D.J.P., AND THORNE, J.A., 1991, Sedimentation on continental Journal of Sedimentary Petrology, v. 63, p. 839–851.
margins, I: a general model for shelf sedimentation, in Swift, D.J.P., WALKER, R.G., AND PLINT, A.G., 1992, Wave- and storm-dominated shallow
Oertel, G.F., Tillman, R.W., and Thorne, J.A., eds., Shelf Sand and marine systems, in Walker, R.G., and James, N.P., eds., Facies Models:
Sandstone Bodies: International Association of Sedimentologists, Response to Sea Level Change: Geological Association of Canada, p.
Special Publication 14, p. 3–31. 219–238.
SWIFT, D.J.P., STANLEY, D.J., AND CURRAY, J.C., 1971, Relict sediments, a WELLS, J.T., AND COLEMAN, J.M., 1981, Physical processes and fine-grained
reconsideration: Journal of Geology v. 79, p. 322–346. sediment dynamics, coast of Surinam, South America: Journal of
SYVITSKI, J.M., HUTTON, E.W., MOREHEAD, M.D., AND COURTNEY, R., 2003, Sedimentary Petrology, v. 51, p. 1069–1075.
Scaling and Integration of Process–Response Stratigraphic Models: WHITLEY, P.K., 1992, The Geology of Heidrun: A giant oil and gas field on
Final Office of Naval Research STRATAFORM report. the Mid-Norwegian Shelf, in Halbouty, M.T., ed., Giant Oil and Gas
TESTA, V., AND BOSENCE, D.W.J., 1999, Physical and biological controls on Fields of the Decade 1978–1988, American Association of Petroleum
the formation of carbonate and siliciclastic bedforms on the north-east Geologists, Memoir 54, p. 383–406.
Brazilian shelf: Sedimentology, v. 46, p. 279–301. WINN, R.D., JR., 1991, Storm deposition in marine sand sheets: Wall Creek
THOMAS, M.A., AND ANDERSON, J., 1994, Sea-level controls on the facies Member, Frontier Formation, Powder River Basin, Wyoming: Journal
architecture of the Trinity/Sabine incised valley system, in Dalrymple, of Sedimentary Petrology, v. 61, p. 86–101.
R.W., Boyd, R., and Zaitlin, B.A., eds., Incised-Valley Systems: Origin WINN, R.D., 1994, Shelf sheet-sand reservoir of the Lower Cretaceous
and Sedimentary Sequences: SEPM, Special Publication 51, p. 85–142. Greensand, North Celtic Sea Basin, offshore Ireland: American Asso-
TILLMAN, R.W., AND MARTINSEN, R.S., 1984, The Shannon shelf ridge sand- ciation of Petroleum Geologists, Bulletin, v. 78, p. 1775–1789.
stone complex, Salt Creek Anticline area, Powder River basin, Wyo- WRIGHT, L.D., FRIEDRICHS, C.T., KIM, S.C., AND SCULLY, M.E., 2001, Effects of
ming, in Tillman, R.W., and Siemers, C.T., eds., Siliciclastic Shelf ambient currents and waves on gravity-driven sediment transport on
Sediments: SEPM, Special Publication 34, p. 1–34. continental shelves: Marine Geology, v. 175, p. 25–45.
TILLMAN, R.W., AND MARTINSEN, R.S., 1987, Sedimentologic characteristics WRIGHT, L.D., WISEMAN, W.J., BORNHOLD, B.D., PRIOR, D.B., SUHAYDA, J.N.,
and production model of Hartzog Draw Field, Wyoming, a Shannon KELLER, G.H., YANG, Z.S., AND FAN, Y.B., 1988, Marine dispersal and
shelf-ridge sandstone, in Tillman, R.W., and Weber, K.J., eds., Reser- deposition of Yellow River silts by gravity driven underflows: Na-
voir Sedimentology: SEPM, Special Publication 40, p. 15–112. ture, v. 332, p. 629–632.
TOEWS, C.N., LECKIE, D.A., AND BERGMAN, K., 2004, Offshore sand ridge YANG, C.S., 1989, Active, moribund, and buried tidal sand ridges in the
accumulation in a low accommodation setting, Bakken Formation, east China Sea and the Southern Yellow Sea: Marine Geology, v. 88,
west-central Saskatchewan (abstract): Canadian Society of Petroleum p. 97–116.
Geologists, Annual Meeting, Proceedings. YANG, C.S., AND SUN, J., 1988, Tidal sand ridges on the East China Sea shelf,
TRAYKOVSKI, P., GEYER, W.R., IRISH, J.D., AND LYNCH, J.F., 2000, The role of in de Boer, P.L., van Gelder, A., and Nio, S.D., eds., Tide-Influenced
wave-induced density-driven fluid mud flows for cross-shelf trans- Sedimentary Environments and Facies: Boston, D. Reidel Publishing
port on the Eel River continental shelf: Continental Shelf Research, v. Company, p. 23–38.
20, p. 2113–2140.
TWICHELL, D.C., 1983, Bedform distribution and inferred sand transport on
Georges Bank, United States Atlantic Continental Shelf: Sedimentol-
ogy, v. 30, p. 695–710.
TYSON, R.V., AND PEARSON, T.H., 1991, Modern and ancient continental
shelf anoxia: an overview, in Tyson, R.V., and Pearson, T.H., eds.,
Modern and Ancient Continental Shelf Anoxia: Geological Society of
London, Special Publication 58, p. 1–24.
TYSON, R.V., AND PEARSON, T.H., eds., 1991, Modern and Ancient Continen-
tal Shelf Anoxia: Geological Society of London, Special Publication 58,
470 p.
VAN DEN MEENE, J.W.H., 1994, The shoreface-connected ridges along the
central Dutch coast: Universiteit Utrecht, Nederlandse Geografische
Studies, v. 174, 222 p.
VAN DER ZWAAN, G.J., AND JORISSEN, F.J., 1991, Biofacial patterns in river-
induced anoxia, in Tyson, R.V., and Pearson, T.H., eds., Modern and
Ancient Continental Shelf Anoxia: Geological Society of London,
Special Publication 58, p. 65-82.
VAN WAGONER, J.C., MITCHUM, R.M., CAMPION, K.M., AND RAHMANIAN, V.D.,
1990, Siliciclastic Sequence Stratigraphy in Well Logs, Cores, and
Outcrops: Concepts for High-Resolution Correlation of Time and
Facies: American Association of Petroleum Geologists, Methods in
Exploration Series, v. 7, 63 p.
VIANNA, A.R., FAUGERES, J.C., AND STOW, D.A.V., 1998, Bottom current-
controlled sand deposits—a review of modern shallow to deep-water
environments: Sedimentary Geology, v. 115, p. 53–80.
398 JOHN R. SUTER
DEEP-WATER TURBIDITES AND SUBMARINE FANS 399

DEEP-WATER TURBIDITES AND SUBMARINE FANS

HENRY W. POSAMENTIER
Anadarko Petroleum Corporation, 1201 Lake Robbins Drive, The Woodlands, Texas 77380, U.S.A.
e-mail: henry_posamentier@anadarko.com
AND
ROGER G. WALKER
Roger Walker Consulting Inc., 83 Scimitar View NW, Calgary, Alberta T3L 2B4, Canada
e-mail: walkerrg@telus.net

Abstract: Depositional environments of deep-water deposits commonly are complex and consequently do not neatly fit any single facies
model. Rather than developing specific models we discuss these deposits within the context of depositional elements and first principles
of process sedimentology. Depositional elements are described using 3D seismic as well as outcrop data. Detailed facies descriptions from
outcrops are then integrated with these depositional elements. Following the theme of this publication, we emphasize facies and
depositional environments rather than the mechanics of turbidity currents and related processes.
The spatial and temporal distribution of depositional elements is determined largely by characteristics of the shelf-edge staging area.
Such factors as grain-size distribution, sediment caliber, frequency of flow events, and magnitude of flows are all a function of conditions
at the shelf edge and upper slope. Sediments are supplied from the staging area to the slope and basin floor beyond. Turbidity currents
traverse the slope through canyons and slope channels. When these flows reach the basin floor they continue to remain confined by levees
for a certain distance. This distance is a function of grain-size distribution in the flow, flow magnitude, and flow velocity. Levee height
diminishes seaward, and eventually where levees can no longer effectively confine the basal sand-rich part of the flow the leveed channel
transitions into a frontal splay or lobe.
Relative sea-level change plays an important role in turbidite deposition, in that sea level is a major factor controlling conditions in
the outer shelf and upper slope. During relative sea-level lowstands, shorelines and consequently depocenters tend to be located at the
shelf edge. This sets up conditions favorable for delivery of sediments to the slope and basin floor. Conversely, relative sea-level
highstands commonly are associated with depocenters at the inner to middle shelf, resulting in a paucity of coarse sediments being
actively delivered to the shelf edge and ultimately to the slope and basin floor. Variations in grain size delivered to the shelf edge during
a cycle of sea-level change can vary predictably hence the temporal and spatial distribution of depositional elements in linked deep-
water environments can likewise be better understood within this context.

INTRODUCTION edition of “Facies Models”. This work will be referenced here but
not repeated in detail.
The scope of turbidite and submarine-fan facies models is
vast, extending from individual beds a few centimeters thick to HISTORY OF FAN MODELS
entire submarine fans with volumes up to a million cubic kilome-
ters or more (for example, Indus fan area 1.1 x 106 km2, thickness The turbidity-current concept was introduced in 1950, in the
3+ km, hence volume of the order of 3 x 106 km3). The unifying classic paper “Turbidity currents as a cause of graded bedding”
theme is the central role played by individual turbidity currents, by Kuenen and Migliorini (1950). The paper was based mainly on
where each bed (a turbidite) is the result of a relatively short-lived Kuenen’s experimental work both before and after the Second
depositional event. The environment is consistently below storm World War. The idea that sand could be transported to great
wave base, such that, once deposited, a turbidite is unlikely to be depths in the ocean was very controversial at the time (Walker,
reworked by other currents aside from the occasional strong 1973), and for many years there was considerable debate about
contour current. Figure 1 schematically illustrates an idealized the existence of turbidites and, about their properties. It was
shelf to basin-floor physiography displaying most of the key understood that modern fans existed, but their internal character-
elements of the deep-water depositional environment. istics were completely unknown—indeed, Kuenen’s experiments
We will briefly examine the history of turbidite and subma- were much more concerned with the origin of submarine canyons
rine fan models and show that, perhaps more than in any other than the transport of sand onto the deep sea floor.
depositional environment, technology (2-D and 3-D seismic data) After a dozen years of observations, the first generalization
has influenced the definition of depositional elements and hence concerning turbidites was published by Bouma, (1962) (Fig. 2)—
the facies models (e.g., Posamentier and Kolla, 2003a). No single what is now known as the “Bouma sequence” for the internal
model comes close to embracing the complexity of huge ancient structures in individual turbidites. The sequence from Division A
and modern submarine-fan systems, making the depositional (generally structureless) to division B (parallel lamination in
elements and their lateral and vertical relationships the basis for sand) and division C (ripple cross-lamination) was compared
interpretation and prediction. Our treatment of facies models will with flume experiments and interpreted to represent waning
take a first-principles approach that will focus on the linkage flow (Harms and Fahnestock, 1965; Walker, 1965). Division D
between physical processes and associated depositional elements. consists of thin laminae of silt and clay, and Division E is pelitic,
Many aspects of turbidity-current generation, movement, probably largely turbidity-current mud with a small proportion
and deposition were reviewed by Walker (1992) in the third of hemipelagic mud

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 397–520.
400 HENRY W. POSAMENTIER AND ROGER G. WALKER

Inner- to Mid-Shelf (Highstand) Delta

Shelf-Edge (Lowstand) Delta

Slump Scars

Canyon/Slope Channel

Staging Area Avulsion Node

Mass-Transport
Complex

Leveed
Channel
Crevasse
Splay

Frontal Splay
Sediment
Waves
Oxbow

FIG. 1.—Schematic representation of shelf to deep-water physiography. The shelf staging area is connected to the deep-water
environment through slope channels and/or canyons. Depositional elements in the deep water include leveed channels, crevasse
splays, sediment waves, and frontal splays or lobes. (modified after Posamentier and Kolla, 2003a).

FIG. 2.—The Bouma (1962) sequence for classical turbidites. Division D is placed in brackets because it is difficult to identify in
weathered or tectonized outcrops. Division E can be subdivided into two parts: turbidite mud E(t) and hemipelagic mud E(h).
In most beds, the turbidite mud predominates.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 401

Gradually, as more information became available on modern (1989) in a classic study of the Mississippi Fan. Weimer (1991)
fans, the first model was proposed by Normark (1970) in another recognized a succession of seismic facies based on reflection
classic paper—“Growth patterns of deep sea fans”. The data base patterns (subparallel, wavy, hummocky, divergent, mounded,
was small, and the paper concentrated on the La Jolla and San and convergent) which could be interpreted in terms of mass-
Lucas fans. The model showed leveed channels on the upper fan, transport complexes (hummocky, mounded), channel fill (dis-
depositional (“suprafan”) lobes on the middle fan, and a smooth continuous, high amplitude), levee (subparallel to convergent),
surface on the lower fan. This model was based largely on and basin floor (subparallel to parallel). Work up to 1991 was
shallow-penetration seismic data. collected in the volume “Seismic Facies and Sedimentary Pro-
Shortly afterward, Emiliano Mutti and colleagues proposed cesses of Submarine Fans and Turbidite Systems” (Weimer and
fan models based exclusively on observations of ancient rocks. Link, 1991). The seismic evidence presented was mostly high-
Mutti and Ricci Lucchi (1972) proposed a model with an inner-fan quality 2-D data. In that volume, Mutti and Normark (1991) first
channel that branched into multiple channels on the mid-fan (but systematized the depositional-elements approach. They defined
without depositional lobes). In the same year, Mutti and Ghibaudo depositional elements as the basic mappable components of both
(1972) showed a similar model but with lobes at the ends of the modern and ancient turbidite systems and stages that can be
channels. Normark’s work was not cited, suggesting that the recognized in marine, outcrop, and subsurface studies. These
modern and ancient fan models were derived independently— features are the building blocks of fan models.
the proposed models suggested that modern fans and ancient The past ten years have seen an explosion in 3-D seismic
rocks behaved in very similar ways. studies, with a corresponding better understanding of deposi-
The channel-feeding-lobe models dominated turbidite stud- tional elements (e.g., Posamentier and Kolla, 2003a). The best
ies for about 10 years (1970–1980). The literature on modern fans sources of this information are in the proceedings volume of a
and ancient rocks was formally brought together into the model Gulf Coast Section of SEPM research conference (“Deep Water
proposed by Walker (1978); the models proved to be popular but Reservoirs of the World”, Weimer et al., 2000) and a thematic
also attracted considerable discussion (e.g., Nilsen, 1980). In compilation of papers on deep-water systems in Marine and
retrospect, the models clearly had severe limitations—the distri- Petroleum Geology (Mutti et al., 2003). Various classifications of
bution of sand and mud on the fans was incorrect, and no depositional elements were suggested, but no attempt was made
consideration was given to the influence of grain size or of local to formulate a general model for submarine fans.
and regional tectonics. Perhaps more importantly, the models did
not incorporate the influences of relative sea-level fluctuation. ORIGIN OF TURBIDITY CURRENTS
With the advent of sequence stratigraphy, the fan models of
the 70s were updated first by Vail et al. (1977) and later by Mutti Density currents flow downslope as gravity acts on the den-
(1985), Posamentier et al. (1988), and Posamentier et al. (1991), sity difference between the flow and the ambient seawater (Fig.
who integrated the effects of relative sea-level fluctuation with 3). The density difference can be due to any or all of the following:
the channel-feeding-lobe models. This can be regarded as a the increased salinity of the flow, the cold temperature of the
period of transition between the older, field-based models and flow, and the suspended sediment within the flow. A turbidity
the rapidly evolving seismic-based models—the technology of current is a special case of a density flow, where the increased
marine geology was overtaking the efforts of field geologists. density is due to sediment maintained in turbulent suspension
In 1982, the first side-scan sonar images of the Amazon Fan within the flow. The turbulence is maintained by the downslope
were published by Damuth et al. (1982a), and Damuth et al. movement of the flow.
(1982b). The presence of long, narrow, and sinuous channels Turbidity currents can originate by two mechanisms: some
surprised most turbidite workers, as did the scale of the channel– begin with large sediment slumps that accelerate and become
levee complexes, which stood at least 200 m above the adjacent turbulent. Many of these slumps are triggered by earthquakes,
fan surface. the most famous being the Grand Banks (Newfoundland)
However, 1985 can be considered the year in which the earthquake, slump, and flow of 1929. The flow broke a series of
emphasis shifted significantly from ancient rocks to large-scale submarine telegraph cables, and reconstructions of the flow
studies of modern fans. In that year, the first compilation of mechanics (Piper et al., 1988) suggest flow velocities up to 20
modern fan studies was published (Bouma et al., 1985), with m/s, flow thicknesses of several hundred meters, and a mini-
discussion of the Amazon, Astoria, Bengal, Cap-Ferrat, Crati, mum flow volume of 175 km3. The flow bypassed the entire
Delgada, Ebro, Indus, La Jolla, Laurentian, Magdalena, Missis- Laurentian Submarine Fan, and the deposit, in places over 1 m
sippi, Monterey, Navy, Rhône, and Wilmington fans. That vol- thick, now covers a large part of the Sohm Abyssal Plain
ume also had a very useful fold-out that tabulated the quantita- (Walker, 1992).
tive descriptors of the fans (channel dimensions and slopes, fan Similarly, slumps off the delta of the Magdalena River in
sizes, and fan volumes). Colombia have broken telegraph cables up to 100 km from the
Also that year, Droz and Bellaiche (1985) published a seismic delta. In the period 1932–1955, there have been 15 cable breaks,
study of the Rhône Fan, showing the existence of meandering averaging one every 1.5 years (Heezen, 1956). The flow of 1935
channels, channel–levee systems, and the lateral shifting and had an estimated volume of sand of 3 x 108 m3 (Heezen, 1956).
stacking of these systems to make channel–levee complexes. Turbidity currents off the fronts of major deltas may be large and
They also showed large slump masses (“acoustically transparent frequent.
units”) up to 160 milliseconds thick that represented both failure In the case of the Congo (Zaire) river, where there is no delta,
on the slope above the fan and failure of the back of the channel a submarine canyon has its head within the estuary of the Congo
levees. River. At times of peak river discharge in December–January and
The studies in the Bouma et al. (1985) compilation essentially April–May, and during the years when the river is establishing a
changed the direction and style of turbidite research, focusing on new course among the estuarine sand bars (1892–1903 and 1925–
modern fans rather than ancient-rock studies. Droz and Bellaiche 1929), submarine cables have been broken seaward of the estuary
(1985), without using the term, essentially introduced the idea of within the Congo Canyon (Heezen et al., 1964). These cables lay
depositional elements. This approach was also applied by Weimer close to the estuary, at the shelf edge, and in water depths as great
402 HENRY W. POSAMENTIER AND ROGER G. WALKER

HORIZONTAL

SLOPE

FIG. 3.—Experimental turbidity current in a flume. Water depth is 28 cm. Note characteristic shape of the head and eddies behind the
head. Sediment is thrown out of the main flow by these eddies, and the body of the flow is about half the height of the head.
Experiment conducted by G.V. Middleton at Caltech.

as 2800 m, suggesting that sand swept into the canyon head from size, sorting, mud content, etc.) were discussed in detail by
the estuary continued down the canyon in flows powerful enough Elmore et al. (1979).
to break cables in abyssal depths. Despite today’s relative high One of the longest documented bypass systems is the Cascadia
stand of sea level, there were 26 cable breaks between 1893 and Deep-Sea Channel (Nelson et al., 2000). The channel originates off
1937—an average of one every 1.7 years. the coast of Washington, continues around the outer part of
Turbidity currents can also originate with delivery of river Astoria Fan, cuts through the Blanco Fracture Zone, and ends on
flow charged with sediment directly onto the slope. During times the Tufts Abyssal plain. The “turbidity-current pathway [traverses]
of river flood enough sediment can be entrained in the flow in 1000 km of Cascadia Basin and remained open throughout the
some instances to produce a mix that has greater density than sea late Quaternary … as shown by the presence of the 13 post-MA
water, resulting in hyperpycnal flow down the slope (Mulder and (Mazama Ash, 7530 YBP) turbidites throughout the pathway in
Syvitski, 1995; Mulder et al., 1998). With this mechanism, what all recent cores we have collected” (Griggs and Kulm, 1968; see
begins as inertial flow at the river mouth transforms into density also Nelson et al., 2000). Beds within the channel include thick (2
underflow and ultimately turbidity flow on the slope. Such flows m) Pleistocene graded gravel-to-sand beds over 400 km from the
generally are of greater duration (i.e., days or weeks) than those heads of the channel at the Washington coast. These long dis-
that originate from large sediment slumps (i.e., hours). tances of bypass have significant implications regarding the
location of sand deposits in ancient basins, as discussed through-
FAN BYPASSING AND DEPOSITION out this review.
ON MODERN ABYSSAL PLAINS Kneller (1995) described the effects of waxing and waning
flows within individual events. Waxing flows, commonly at or
Turbidity currents traveling downslope may be moving at near the head of a turbulent flow, erode the substrate over which
several to many meters per second, at which velocities all of the they pass. Significant amounts of sediment can bypass the system
sand and finer sizes are in turbulent suspension. The flows during this time. As the flow wanes, coarser sediments tend to
gradually decelerate to velocities of 1–2 m per second, when the come out of suspension and be deposited in the area formerly
coarser sand fraction begins to be deposited from suspension. characterized as a zone of bypass. Consequently, even in slope
During this period of deceleration, the flows may largely bypass and proximal basin-floor areas, where sediment bypass and
the slope and move long distances across the basin floor. erosion during waxing phase may be common, some sedimenta-
There are many studies of abyssal-plain deposition (Pilkey, tion in the form of lag deposits almost always occurs.
1988). The Grand Banks flow bypassed the Laurentian Fan at the
base of the slope and deposited a turbidite on the Sohm Abyssal TURBIDITE FACIES—THE BUILDING BLOCKS
Plain, as discussed above. About 16,000 years ago, an even
larger flow bypassed the Hatteras Fan and deposited on the There are several schemes for classifying the family of rocks
Hatteras Abyssal Plain (western North Atlantic Ocean) (Elmore that occur in deep-water settings. The first was proposed by Mutti
et al., 1979). Deposition began about 120 km from the end of the and Ricchi Lucchi (1972) and was later simplified by Walker in
Hatteras Canyon system. This “Black Shell turbidite” (named 1978. Subsequent facies classifications have become more com-
for the distinctive corroded shells contained in the deposit) plex, including the all-inclusive but unwieldy schemes of
covers 44,000 km2 of Hatteras Abyssal Plain in a bed up to 4 m Ghibaudo and Vanz (1987) and Pickering et al. (1986).
thick, 500 km long, 200 km wide. The volume of the deposit is A detailed subdivision of features within individual beds was
between 100 and 200 km3. Characteristics of that deposit (grain proposed by Lowe (1982), based on interpretations of how sedi-
DEEP-WATER TURBIDITES AND SUBMARINE FANS 403

ment was deposited from sandy and gravelly high-density tur- when in nature flows may change and evolve over distances of
bidity currents. For sandy flows, division S1 is characterized by hundreds of kilometers).
traction structures, division S2 contains “thin horizontal layers
showing inverse grading and basal shear laminations” and divi- Classical Turbidites
sion S3 “may be structureless or normally graded and it com-
monly contains water escape features”. For gravelly flows, divi- This category includes all of those rocks originally considered
sion R1 consists of coarse gravel with traction structures, division as turbidites in the 1950s and 1960s—the beds that give rise to
R2 consists of an inversely graded gravel layer, and division R3 little or no controversy today. The facies includes thick monoto-
consists of a normally graded gravel layer. Lowe’s (1982) scheme nous successions of alternating sandstones and mudstones (Fig.
is akin to a Bouma sequence (see below) for individual coarse 4). The sandstones have sharp, flat bases, and the only erosional
beds, rather than a facies classification of coarse-grained beds. features are normally on the centimeter scale. They include scour
Because it is genetically based, the scheme may change as more is marks (commonly flute casts) and tool marks (commonly groove
learned about the flow and depositional mechanics of high- casts). Channeling on a scale greater than a meter is very uncom-
density flow events. mon.
In this review, we suggest that deep-water rocks contain a Internally, classical turbidites contain some or all of the
variety of depositional elements (discussed below) and that these divisions first proposed by Bouma (1962) (Fig. 2). Division A
elements contain distinctive assemblages of facies. We have implies rapid deposition producing structureless sandstone in
chosen to use the simple scheme of Walker (1978), which is the absence of any equilibrium bedforms, whereas divisions B
descriptive (except for the various deformed facies), and based on and C imply traction of grains on the bed to form parallel
grain size. The categories included in this scheme are (1) classical lamination and ripple cross lamination, respectively (a waning-
turbidites, (2) structureless sandstones, (3) pebbly sandstones, (4) flow succession—Harms and Fahnestock, 1965; Walker, 1965).
conglomerates, and (5) various types of deformed rocks. We are Divisions D and E both imply deposition of fine-grained mate-
more concerned with the descriptive and environmental aspect rial from suspension without traction on the bed. Note that
of the facies than with the mechanics of flow and deposition Bouma (1962) observed that division D (laminations of silt and
(which are very difficult to study in flumes a few meters long mud) was difficult to recognize in “weathered or tectonized

TOP

FIG. 4.—Alternating beds of sandstone and mudstone, Devonian, Cape Liptrap, South Australia. Note the monotonous alternation
of sandstones and mudstones, and the very parallel nature of the bedding with no evidence of any topography on the sea floor.
404 HENRY W. POSAMENTIER AND ROGER G. WALKER

outcrops”—consequently it may not be a useful or significant 8A; Walker, 1985); C for climbing, C for convolution and C for
part of the Bouma sequence. clasts.
Two sub-categories of classical turbidites have been sug- The presence of convolute lamination implies rapid deposi-
gested by several workers: thin-bedded and thick-bedded. It tion of sediment and trapping of pore fluid, such that the primary
must be emphasized that there is a complete spectrum of bed structures are easily deformed. The climbing ripples imply depo-
thicknesses and that their separation is arbitrary. The thick- sition of sediment from suspension while the ripples are moving
bedded turbidites (sandstones roughly in the range 10–100 cm in on the bed. Within the category of thin-bedded turbidites, the
thickness; Fig. 5) tend to be composed of Bouma’s division A, simple beds imply traction on the bed and essentially no deposi-
with fewer beds also containing divisions B and C (Walker, 1968). tion from suspension, whereas the CCC turbidites suggest high
Thin-bedded turbidites (Fig. 4) tend to lack Bouma’s division A, rates of deposition from suspension during formation of the
and the sandstones contain only the B–C or C divisions. The primary bedforms. It has been suggested that thin beds showing
nature of these divisions suggests two distinct types of thin- high rates of deposition commonly form on levees. The thin beds
bedded turbidites. that show little evidence of rapid deposition from suspension
The simplest type of thin-bedded turbidite contains a single may indicate basin-plain settings, where the turbidity currents
set of ripple cross lamination, with or without division B parallel have much less sediment left in suspension (Fig. 8B; Walker,
lamination underneath (Fig. 6). In more complex thin-bedded 1985). The presence of ripped-up mudstone clasts supports this
turbidites, the ripples in division C consist of climbing sets rather interpretation—there is more likelihood of erosion associated
than single sets, and the ripple cross lamination (and the parallel with confined flows in channels than in unconfined settings on a
lamination beneath) can be convoluted (Fig. 7). These complex distal basin plain.
thin-bedded turbidites also commonly contain ripped-up mud-
stone clasts, and they have been termed “CCC turbidites” (Fig. Structureless Sandstones
There is an association in facies between classical turbidites
and structureless sandstones. Individual structureless sandstone
beds tend to be thicker (several tens of centimeters to a few
meters) than the sandstones in classical turbidites, and mudstone
partings between beds tend to be thin (centimeters) or absent (Fig.
5). The deposits of several flows may be amalgamated, the
amalgamation planes being denoted by (1) abrupt changes in
grain size, (2) layers of ripped-up mudstone clasts, or (3) simply
the disappearance of thin mudstone partings (Fig. 9). On a larger
scale, scouring on the scale of meters is commonly observed in
this facies (Fig. 9). It follows that the monotonous interbedding of
sandstones and mudstones, typical of classical turbidites, does
not occur in structureless sandstones. Stacks of amalgamated
beds without mudstone partings can be as much as 200 m thick,
as in the Annot Sandstone (Fig. 5).
Parallel lamination and ripple cross lamination are rare, and
the term structureless (now preferred to the older term “mas-
sive”) denotes this absence of primary sedimentary structures.
Graded bedding is present in some beds and not in others; its
presence may be largely a function of the range of grain sizes
available in the flow.
Although most beds lack primary structures, secondary
structures indicating dewatering during compaction of the bed
are common (Lowe, 1975). These include vertical or subvertical
fluid-escape pipes (Fig. 10), which can become contorted if the
bed is sheared by continuing turbidity-current flow during the
fluid escape (Fig. 11). If the escaping water encounters a crude,
incipient parallel lamination with variations in permeability,
the water may be forced to flow horizontally until able to break
through the less permeable layers and continue its vertical
escape. The curved upward edges of these laminates take the
shape of an irregular stack of dishes, hence the term “dish
structure” (Figs. 10, 12).
The association of this facies with classical turbidites suggests
that individual structureless sandstones are also the deposits of
turbidity currents. This interpretation is strengthened by the
presence of fluid-escape features, which indicate initial deposi-
tion of a fluid-rich sediment–water mixture (rather than a more
rigid plug flow with grain-to-grain contacts and much less inter-
FIG. 5.—Thick-bedded sandstones consisting mainly of Bouma’s stitial water). Despite the thickness of individual beds and the
division A, separated by very thin siltstone partings. Height general absence of Bouma sequences, there is no compelling
of cliff about 180 m. Compare with Figure 4. Annot Sandstone observational or experimental evidence to reject turbidity cur-
(Eocene), southern France. rents in favor of speculative processes such as fluxoturbidity
DEEP-WATER TURBIDITES AND SUBMARINE FANS 405

FIG. 6.—Thin-bedded turbidites beginning with Bouma divisions B and C. Sharp bases shown by yellow arrows, parallel lamination
by a blue arrow, and ripple cross lamination by red arrows. Note the absence of convolute lamination, climbing ripples, and
ripped-up mudstone clasts. Ordovician turbidites at Chutes Montmorency, Quebec.

CONVOLUTE
LAMINATION

CLIMBING RIPPLES

FIG. 7.—Thin-bedded turbidites in the Chatsworth Sandstone (Cretaceous), Chatsworth (Simi Hills), California. Bases shown by red
arrows, and climbing ripples shown by yellow arrows. Convolute lamination is outlined in blue. Compare with Figure 6 (where
there is no climbing and no convolution).
406 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 8.—Diagram showing two types of thin-bedded turbidites. One is characterized by single rows of ripple cross lamination without
climbing, and the other is characterized by climbing ripples, convolute lamination, and ripped-up mudstone clasts A). The “CCC
turbidites” are interpreted as levee deposits (see text), and the others as distal basin plain deposits B) (From Walker, 1985).

TOP

STRUCTURELESS SANDSTONES

CLASSICAL
TURBIDITES

SCOUR

AMALGAMATION

FIG. 9.—Devonian turbidites in Germany. Note the classical turbidites (right) and the underlying thick-bedded structureless
sandstones. Yellow arrows show thin mudstone partings that disappear along strike (amalgamation), and the red arrow shows
a small scour.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 407

A)
A) DISH
DISH STRUCTURES
STRUCTURES

VERTICAL
VERTICAL PIPES
PIPES

TOP

FIG. 10.—Vertical fluid-escape pipes with overlying dish structures (outlined in yellow).
Ordovician Cap Enrage Formation, Gaspesie, Quebec.

currents (Dzulynski et al., 1959) or sandy debris flows


(Shanmugam et al., 1994; Shanmugam, 1996).

Pebbly Sandstones F IG . 11.—Distorted


fluid-escape pipes
As the coarse fraction within flows gradually increases, the in a core from the
structureless sandstone facies grades through granule sandstones Cretaceous Lysing
into the pebbly sandstone facies. Graded bedding is common (Fig. Formation, off-
13) and readily observed because of the wide range of sizes present. shore mid-Nor-
Internally, beds may show a crude horizontal stratification, and, in way. Well 6506/
rare cases, planar tabular and trough cross bedding may be present. 12-4, 3240.6 m
In the Cap Enrage Formation (Quebec), the trough sets are up to 50 depth. 5 cm
cm thick, and trough widths seen in plan view are up to at least two
meters. Apart from structureless sandstone, the elements of the
Bouma sequence do not occur in pebbly sandstones, and hence the
Bouma sequence cannot be used as a descriptor in this facies
If blade-shaped or disc-shaped pebbles are present, they are
commonly well imbricated (Fig. 14). The features described
above—graded bedding, cross bedding, and imbrication—all sedimentary record. Four distinct facies were recognized by
suggest turbulent flows in which grains are free to move relative Walker (1975a), but the classification is based on a relatively small
to one another, enabling the development of these features. sample and does not have the authority of the Bouma sequence
Making reasonable estimates of the turbulence of the flow and for classical turbidites. The features used to define the facies are
particle settling velocities, it appears that a flow moving at 6 m/ (1) the style of grading (normal or inverse), (2) the type of
s (the Grand Banks flow of 1929 near the toe of Laurentian Fan; stratification, and (3) the fabric. In combination, these features
Uchupi and Austin, 1979) could suspend by fluid turbulence define the four facies.
alone clasts up to 2 or 3 cm in diameter. It therefore appears that The first consists of beds which are normally graded and pass
pebbles can be transported into deep water by turbidity currents upward into finer-grained stratified pebbly sandstone (Fig. 15).
(flow velocities of 6 m/s or greater), and that such flows could The second consists of beds that show only normal grading (Fig.
deposit graded, imbricated, and/or cross-bedded beds. In these 16) without a stratified component. The third consists of beds that
instances, despite the coarse nature of the beds, it is again not begin with inverse grading and pass upward into normally
necessary to appeal to alternative transport processes such as graded beds (Fig. 17). Finally, the fourth facies lacks any of these
fluxoturbidity currents and sandy debris flows. features and is described as disorganized or structureless (Fig.
18).
Conglomerates The first three facies may also display clast imbrication (Figs.
14, 17). In the stratified parts of the graded-stratified facies,
Conglomerates are not as common as the facies described clasts lie with their long axes transverse to flow, and the short
above, but they do make up an important part of the deep-water axis dips upstream. In the graded and inversely graded parts of
408 HENRY W. POSAMENTIER AND ROGER G. WALKER

3255.25 m 3255.44 m

FIG. 12.—Dish structures out-


DISH
DISH STRUCTURE
STRUCTURE lined in yellow from the
Agat 35/3-4well, 3255.5
m, offshore Norway. Note
darker (less permeable)
layer at base of each dish,
and a few fluid-escape
pipes red arrows) where
fluid has broken through
vertically.

AGAT. 35/3-4

5 cm

the beds, however, clasts more commonly lie with their long
axes parallel to flow, with the long axes dipping upstream. This
fabric suggests that the clasts have not been rolling on the bed
(if clasts roll, long axes tend to be transverse to flow). A full
discussion of conglomerate fabrics has been given by Walker
(1975a).
Bed thicknesses in the conglomerate facies are very variable.
Individual graded beds can be over 10 m thick (Fig. 16), but
alternatively, beds may be only one or two pebble diameters in
thickness. The example from Point Lobos, California, in Figure
19, shows several thin conglomerate layers alternating with
sandstone layers. The pebbles probably did not constitute the
bulk of the flow, because they would not have given a sufficient
density contrast with the surrounding fluid to drive the flow. It
is more likely that the flows that transported the pebbles were
large sandy turbidity currents, and that the thin pebble beds
represent lags left behind by the main flows.
Despite the suggestions made above, interpretations of con-
glomerate facies remain somewhat speculative because of the
lack of large-scale experimental work.

EXOTIC FACIES—OTHER TYPES OF DEPOSITS


IN DEEP-WATER ENVIRONMENTS

This category contains a variety of facies that do not fit into the
four facies described above. They are generally characterized by
poor sorting and lack of coherent bedding features. Some of the
main types are described below.


FIG. 13.—Pebbly sandstone about 1 m thick showing overall
graded bedding. Annot Sandstone at Chambre du Roi, south-
ern France. Fixed eyebolts for rock climbers are circled for
scale.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 409

FLOW

F IG . 14.—Graded con-
glomerate showing
clast imbrication (cen-
ter of bed above note-
book). Bed rests on de-
formed and slumped
mudstones. Tourma-
line State Surfing
Beach, north of San
Diego, California.

Pebbly Mudstones

Pebbly mudstones (Fig. 20) consist of granules and pebbles,


commonly along with distorted sandstone clasts, all embedded
in a deformed mudstone matrix. The term was introduced by
Crowell (1957), who suggested two mechanisms for their origin.
The first was emplacement by debris flows, wherein the strength
of the muddy matrix prevents the settling of the larger pebbles
and clasts during transport. Some debris flows may slide rapidly
on a basal layer of lubricating fluid (sea water) (Mohrig et al.,
1988), though it is likely that at least part of the moving mass
clearly is in contact with and erodes the substrate over which they
pass (Fig. 21; Posamentier and Kolla, 2003a) suggesting a more
complex rheology. As the flow velocity increases, there is a
tendency for sediment to be suspended at the head as well as the
upper parts of the flow. Rapidly moving debris flows therefore
may tend to transform at least in part into turbidity currents. The
transformation may be quite slow for muddy flows, but for sandy
debris flows moving at more than 1 to 2 m/s (the velocity at which
sand is carried in suspension) the transformation into a turbulent
turbidity current may be rapid, and take place over a short
distance.
A second mechanism for depositing pebbly mudstones
(Crowell, 1957) involves the passage of a sandy/pebbly turbidity
current over a bed of fluid-saturated, uncompacted mud. The
coarser material from the flow may be deposited on the muddy
surface and then quickly sinks into the uncompacted mud. The
pebble–sand–mud mixture may then flow for a short distance as
it dewaters, mixing the various grain sizes and then depositing an
ungraded, poorly sorted pebbly mudstone.


FIG. 15.—Graded conglomerate in the Cap Enrage Formation,
Ordovician, Quebec. Note the large carbonate blocks in the
base of the bed, and the gradation into structureless pebbly
sandstone, stratified pebbly sandstone, and finally structure-
less sandstone. Bed is at least 8 m thick.
410 HENRY W. POSAMENTIER AND ROGER G. WALKER

BEDDING
BEDDING

TOP

FIG. 16.—Graded conglomerate about 14 m thick, Eocene of Oregon. Close field examination showed a progressive decrease in
maximum and estimated mean grain size throughout the bed, and no internal planes or grain size changes that might suggest
an amalgamated bed.

INVERSE
INVERSE GRADED
GRADED
BEDDING
BEDDING

BASE
BASE

FIG. 17.—Inversely to normally graded conglomerate, Cretaceous La Jolla Formation, California. Note also the well developed
imbrication with clasts dipping upstream (to the right).
DEEP-WATER TURBIDITES AND SUBMARINE FANS 411

TOP

STRUCTURELESS

FIG. 18.—Conglomerate in the Ordovician Cap Enrage Formation, Quebec, showing no grading and no imbrication. The bed is
described as disorganized or structureless.

SANDSTONE
SANDSTONE LAYERS
LAYERS

TOP

FIG. 19.—Thin conglomerate horizons separated by sandstone layers (shown by yellow arrows). Cretaceous, Point Lobos,
California.
412 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 20.—Pebbly mudstones from Pigeon Point, California. This is the classic location where Crowell (1957) first discussed the
origin of pebbly mudstones. Note clasts scattered throughout the muddy matrix, along with rolled-up sandstone beds (above
notebook).

Slumps Slumps with Stratified Blocks

Slumps (Figs. 22, 23) comprise a large category of variously Slumps with stratified blocks (Fig. 24) are not uncommon. The
deformed sediments. The term in its most general sense de- stratified blocks may be meters in diameter, and they consist of
scribes chaotic unbedded units meters to tens of meters in interbedded layers of sandstone and mudstone (perhaps origi-
thickness. The lithology may consist only of mudstone (Fig. nally classical turbidites). In most cases the blocks were probably
22), or it may involve pulled-apart or rolled-up sandstone beds not lithified, so that transport in a turbulent medium would
in a matrix of mudstone. In all cases, deposition was probably probably result in splitting of the blocks along the cohesionless
fairly rapid, and mudstones form an important part of the sandstone layers. This would destroy the stratified blocks and
facies. Transport distances vary from a few meters (e.g., col- probably would give rise to large mudstone clasts. One sugges-
lapse of channel walls) to hundreds of kilometers across basin tion is that the blocks were derived from an undermined, col-
floors. Original depositional conditions may have involved lapsed channel wall, where the blocks subsequently were buried
rapid deposition of sandstone with trapped pore water, fol- by turbidity-current sediment before they could be transported
lowed by mudstone deposition and sealing of the pore fluid. downchannel. Later in this paper, examples will be shown of
With continued deposition, the lithostatic load would in- large blocks that have been rafted on top of mass-transport
crease, but, if the pore fluid could not escape, the fluid pressure complexes. Such flows are not turbulent, and the rafted blocks
would also increase. Slumping would be initiated if beds fail may retain some stratification.
along a weak layer with high pore pressure. The sediment may Slumps involving only one or two beds (Fig. 25) are fairly
move a few meters during dewatering and deposit a unit that common. The deposit is characterized by undeformed bedding
consequently would be identified as a slump. Alternatively, if below and above the slumped horizon, and coherent but rolled-
the sediment moved a greater distance the resulting deposit up beds within the slump. The beds are commonly thin (a few
may be identified as a debris flow. Some slumps may move fast tens of centimeters maximum), and are associated with other thin
enough so that much of the sediment is taken into suspension beds and with “CCC” turbidites (Walker, 1985) interpreted as
with turbulence characterizing the flow, transforming the levee deposits. Thus the slumps may indicate rapid deposition of
mass into a turbidity current. beds on the levee (either the side facing the channel, though more
DEEP-WATER TURBIDITES AND SUBMARINE FANS 413

A N

Flow Direction
1.5 km

one km

S 50 m N
S
B
one km

50 msec

one km

SSW NNE

FIG. 21.—A) Plan view of the erosional base of a mass-transport complex in the ultra-deep environment of the Makassar Strait,
Indonesia. Parallel to divergent erosional grooves are observed. B) Section-view image of mass-transport complex. The mass-
transport complex is characterized by chaotic seismic reflection character, with erosional scour in excess of 50 m locally, and up
to 1.5 km wide (modified from Posamentier et al., 2000). Seismic data courtesy of WesternGeco.

FIG. 22.—Large slump resulting in almost complete disruption of bedding, Carboniferous Bude Sandstones at Efford, southwest
England.
414 HENRY W. POSAMENTIER AND ROGER G. WALKER

commonly on the back side of the levee), perhaps with trapped


pore fluid within the sandstones. The slopes associated with the
levee must have been sufficient to allow gentle sliding of just one
or two beds, without large-scale deformation of the underlying
sediment

CONTROLLING FACTORS
ON DEEP-WATER SYSTEMS

Here, we discuss some of the principles relevant to the depo-


sition of turbidites and the depositional elements within which
they occur. We will “set the stage” with regard to the context
within which deep-water depositional elements are deposited. A
sound understanding of process is key to the construction of
robust depositional models, enabling geoscientists to construct
models that will be applicable to their unique set of environmen-
tal circumstances. Such models can then be a useful predictor of


FIG. 23.—Slump folds in Eocene slope mudstones of the Cozy Dell
Formation, Highway 33 north of Wheeler Gorge, California.
Scale shown by notebook.

FIG. 24 (below).—Slump involving large stratified blocks in the


Upper Cretaceous Great Valley Sequence, Lake Berryessa,
California. The two heavy black lines show bedding (top to
the left)—the bedding is parallel but the lines converge be-
cause the camera is pointed steeply up the cliff. The matrix is
a silty mudstone with a large variety of pebbles and cobbles.
The stratified blocks consist of layers of sandstone and mud-
stone. It is argued that these would easily be disintegrated
along the unconsolidated sandstone layers if there had been
significant transport. It follows that they may have collapsed
from a nearby channel wall and were buried before they could
break up. Slumped bed is about 7 m thick.

BEDDING
DEEP-WATER TURBIDITES AND SUBMARINE FANS 415

BEDDING

15
15 cm
cm

FIG. 25.—Small-scale slump involving only two beds within otherwise flat-bedded succession, Eocene, Waitemata Group, New
Zealand.

spatial and temporal lithofacies distribution. A typical deep- style of turbidite deposition downslope, a relationship described
water depositional environment with associated depositional by Reading and Richards (1994). Posamentier and Kolla (2003a)
elements is illustrated in Figure 1. discuss how grain-size distribution exerts this control on the style
of turbidite deposition, and is schematically illustrated in Figure
Sediment Staging Areas 26. This figure illustrates the relationship between total flow
height, the height of the high-density part of the flow, and levee
The staging area can be defined as the shelf and/or upper- height, and the resulting transition between leveed channel and
slope location where turbulent flows originate (Fig. 1). This frontal splay in the absence of a change in the gradient of the
staging area and the characteristics of the sediments delivered to slope.
that area are all-important in dictating the nature of turbidity As flows travel down-system, they progressively become
currents and subsequently their deposits farther down-system. better organized, with finer sediments concentrating in the upper
In particular, the sand-to-mud ratio that characterizes these sedi- part of the flow and coarser sediments concentrating in the lower
ments plays an important role in determining whether long part of the flow. The result is that the upper part of the flow tends
leveed channels will develop down-system or whether short to have a lower density and concentration than the lower part of
leveed channels feeding frontal splays or lobes will characterize the flow. The tops of many turbulent flows are higher than the
downslope areas instead. The sediments that ultimately get in- associated levee crests (Fig. 27), the result of which is that the
corporated into flows can be delivered to the staging area by lower part of the flow, where much of the sand-size sediment is
fluvial, eolian, or longshore-drift processes. Subsequent turbidity concentrated, is fully confined by the channel walls. In contrast,
currents originate as sediment failures, associated with seismic- the upper part of the flow, which rides well above the levee crests,
ity and slope instability. Alternatively, if rivers deliver sediment is largely unconfined by channel walls and hence is free to expand
directly to canyon heads, high-density flows within rivers can laterally beyond the levee crests and onto the overbank. This
continue directly into the deeper basin by density underflow (i.e., process of flow spillover results in the deposition of thin, fine-
hyperpycnal flow). Such density underflows can transform into grained turbidites (commonly CCC turbidites; Figs. 7, 8) on the
true gravity flows farther down the slope. crests and backs of the levees, and it also results in progressive
impoverishment of mud within the total flow. In addition to mud
The Significance of Sand-to-Mud Ratio within Flows being lost from the flow by spillover, some sand may also be lost
from the flow because of sedimentation at the flow base and by
The initial sand-to-mud ratio within flows is dictated largely mixing with the remaining upper, lower-density part of the flow.
by conditions in the staging area. The grain-size distribution in However, the amount of sand lost from the flow is volumetrically
these shelf-edge depocenters ultimately plays a critical role in the significantly less than the amount of mud lost from the flow by
416 HENRY W. POSAMENTIER AND ROGER G. WALKER

Sand:Mud
Total Flow Height High
(i.e., Height of low-density +
high density columns)

Levee Height

Potential Overbank
Height Sediment Supply

Channel Complex
Single Leveed

(Frontal Splay)
Distributary
Channel
Low
Proximal Distance Down-System Distal
“Effective” Flow Height
Transition Point
(i.e., Height of high-density
column within turbidity flow)

FIG. 26.—Schematic depiction of the interplay between sediment gravity flows, net sand, and levee height with distance down-
system. Note that the high-density part of the gravity flow is located progressively more closely to the levee crest with distance
seaward. A transition from leveed channel to frontal splay/lobe occurs when the high-density part of the flow (i.e., the sand-rich
part of the flow) reaches bankfull stage. Note also that the highest sand-to-mud ratio occurs there as well (modified from
Posamentier and Kolla, 2003a).

Levee crests

Potentially prospective
overbank deposit

Confined channel flow

Overspill

Flow stripping

FIG. 27.—Schematic illustration of sediment gravity flow through a leveed channel (compare with Fig. 77). The cross-sectional view
illustrates that the flow top lies well above the levee crest. The part of the flow between the flow top and levee crest is unconfined
and systematically spills out of the channel. Enhanced spillover occurs at outer bends (by the process of flowstripping). These
locations constitute areas of preferred sand deposition in the levee environment.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 417

spillover. Thus the channel floor tends to aggrade somewhat of the transition point (Figs. 29, 30). Thus, if a succession of sand-
more slowly than the levee crests. rich flows is followed by a succession of mud-rich flows, the
Also, somewhat counterintuitively, flows tend to become transition point shifts seaward and the result is the superposi-
sandier down-system (i.e., flows have a progressively higher tion of a single leveed channel across an older frontal splay (Fig.
sand-to-mud ratio down-system) as a result of continual prefer- 31). Under certain circumstances, where later flows are more
ential shedding of muddier sediment due to spillover. Kolla and sand rich, the reverse can occur as well. As previously dis-
Coumes (1987), Pirmez and Flood (1997), and Hiscott et al. (1997) cussed, such changes in sand-to-mud ratio commonly originate
have observed that with increased distance down-system there is in the staging area, and are manifestations of changing propor-
a gradual increase of net sand deposited on levees, consistent tions and rates of delivery by rivers and shelf processes, of
with a progressive impoverishment of mud within flows in the different sediment sizes.
down-system direction. The progressive loss of the upper or
mud-rich part of the flow results in a progressive decrease in Slope and Basin Physiography
levee height down-system (Fig. 28). At some point down-system,
the high-concentration or sand-rich part of the flow reaches the The morphology of the slope and basin floor influences the
levee crests (bankfull stage). This is a critical location because deposition of turbidites in a variety of ways. Physiographic
down-system from this point spillover is no longer associated factors include (1) sea-floor rugosity on a large scale such as fault
mainly with the muddy part of the flow; rather, sand-rich flows scarps and intraslope basins associated with salt tectonics or toe-
are now directed across the overbank. Geomorphologically this is of-slope thrust faults, (2) small-scale sea-floor rugosity compris-
expressed as a transition from a single leveed channel to a ing sea-floor irregularities associated with earlier depositional
distributary channel complex or frontal splay (Posamentier and events such as slides or debris flows, (3) the height of the available
Kolla, 2003a). This location, referred to as the transition point, also relief from shelf edge to basin floor, (4) the gradient of the slope,
marks the location where the sand-to-mud ratio within the flow (5) the presence of significant breaks in slope such as those that
is greatest. Downslope of this location the rate of sand being lost can occur where the slope meets the basin floor, and 6) the rate of
from the flow exceeds the rate of mud being lost from the flow, change of the slope.
largely because the sand-rich part of the flow is now largely Perhaps the most well documented example of the effect of
unconfined. The increased cross-sectional area of the floor results sea-floor rugosity on turbidite systems is the fill-and-spill model
in decreased flow velocity and sand deposition. of turbidite systems that characterizes salt-supported intraslope
The sand-to-mud ratio in the flow is critical to this analysis basins (Prather et al., 1998). They described a scenario whereby a
insofar as changing this flow characteristic changes the location string of intraslope basins would fill progressively from the

West East
FIG. 28.—Arbitrary seismic section constructed along a levee
crest. Levee facies are characterized by low-amplitude dis-
continuous reflections and overlie a high-amplitude continu-
ous to discontinuous frontal-splay complex (see Figs. 64 and
116). The section, which is flattened on the levee top, shows a
n
progressive decrease in levee thickness from landward to wD
low
Flo
F Dirireecctitioon
seaward. It also shows a consistently thicker levee along outer X’
bends than along inner bends. Most notably, the location
where the levee thickness approaches zero is also where the
confined flow within the leveed channel transitions into a X
frontal splay/lobe. This location is referred to as the transition
point. Seismic data courtesy of WesternGeco.

Transition Point
West Decreasing levee height East
Top Levee

100 msec

Base Levee

Inside Bend

five km
Outside Bend
Seaward
418 HENRY W. POSAMENTIER AND ROGER G. WALKER

A Transition Point
Time 1
Leveed channel
Lev
ee C High Sand:Mud
rest
Height
(Early Lowstand)
Flow Top

Top of sand-rich
part of flow

Distance Down-System

Seaward Shift of Transition Point


Frontal splay

Time 2 Transition Point

Decreasing Sand:Mud
B
Leve Transition Point
e Cr
Height

est
Flow Top

Top of sand-rich
part of flow Distance Down-System

Time 3
Flow Top Low Sand:Mud
C Leve
(Late Lowstand)
e Cr Transition Point
est
Height

Top of sand-rich
part of flow Distance Down-System

FIG. 29.—Shift of the transition point in response to differences in sand-to-mud ratio with sediment gravity flows. A) A high sand-
to-mud ratio is associated with a transition point that is significantly farther landward than is the case with a lower sand-to-mud
ratio B, C).

Transition Point Transition Point


Time of high sand:mud Time of low sand:mud

A'

Migration of transition point in response


to change from high to low sand:mud

Leveed channel

A A'

Frontal splay

FIG. 30.—Superposition of leveed-channel system over a frontal splay/lobe, which would accompany a progressive muddying-up
of successive sediment gravity flows (compare with Figs. 31, 74, 116, and 163).
DEEP-WATER TURBIDITES AND SUBMARINE FANS 419

5 km 5 km

Late
Late Lowstand
Lowstand
Late
Late Lowstand
Lowstand

Early
Early Lowstand
Lowstand Early
Early Lowstand
Lowstand
Late Lowstand

Early Lowstand
FIG. 31.—Two examples of the superposition of late lowstand leveed-channel deposits over early lowstand frontal-splay deposits
(compare with Fig. 30). Seismic data courtesy of WesternGeco.

proximal basin to ever more distal basins through time. Within The effect of a slope to intraslope-basin transition is shown in
each intraslope basin, turbidite deposition comprises a succes- Figure 32B and C. Flow vectors on the slope are directed primarily
sion from sheet-bedded deposits at the base to leveed-channel downslope parallel to flow. Upon encountering an intraslope
deposits near the top. Flows entering an intraslope basin would basin, laterally directed flow vectors are significantly enhanced.
initially encounter markedly concave-up topography. This con- This increases the likelihood of levee breaching and a resultant
cave-up morphology results in a rapid deceleration of flow, distributive channel pattern and deposition of a frontal splay.
which in turn favors deposition of sheet-like frontal splay depos- Sea-floor rugosity associated with fault scarps or other abrupt
its (Posamentier and Kolla, 2003a). As the basin gradually fills, changes in slope gradient can also have profound influence on
the topography progressively becomes less concave up. The turbidite deposition. In general, the farther a flow travels without
response is an upward transition from frontal splays to leveed any significant breaks in slope the more the sand grains tend to
channels. As a result, leveed-channel deposits tend to dominate become concentrated towards the base of the flow and mud tends
the upper part of intraslope basin fill. This basin-fill evolution to concentrate towards the top. Thus grain-size sorting or segre-
from frontal-splay dominated to leveed-channel dominated is an gation within the flow progressively improves down-system—
autocyclic phenomenon occurring in response to the evolution of the flow becomes better organized. The presence of an abrupt
local topography. Posamentier and Kolla (2003a) describe a ma- slope change in the path of the flow results in a perturbation (i.e.,
trix of possible outcomes associated with varying slope concavity a hydraulic jump; Komar, 1971) within the flow and a consequent
and varying sand-to-mud ratio (Fig. 32A). tendency for abrupt increased flow disorganization. This pertur-
420 HENRY W. POSAMENTIER AND ROGER G. WALKER

C Slo
pe/
bas
B in
flo
or
tra
ns
itio
n

Time 1

Time 2

Time 3
FIG. 32.—A) Matrix of possible responses of location of transition point to varying sand-to-mud ratio and varying slope curvature.
Increased slope curvature, a local or autocyclic parameter, results in a landward shift of the transition point. Increased sand-to-
mud ratio, an external or allocyclic parameter, has a similar effect (after Posamentier and Kolla, 2003a). B) Turbidity flow through
a leveed channel onto a basin floor. C) Upon encountering the basin floor, laterally directed flow vectors are significantly and
abruptly increased. This can result in deposition of a frontal splay on the basin floor. The degree to which a frontal splay forms
at the transition from slope to basin floor is a function of the abruptness of the slope change at this location.

bation results in poorer grain-size sorting within the body of the with a very different set of turbidite depositional elements on the
flow down-system of the abrupt slope change. Consequently, associated basin floor. As illustrated in Figure 34A the flow that
settings such as the base of a fault scarp or an abrupt slope-to- has reached the basin floor early in its run has not had the chance
basin transition can cause significant reorganization of flow. The to become organized from a grain-size distribution perspective,
result of this reorganization is the tendency for the system to hence the transition point is located significantly farther land-
change from relatively confined leveed channels to relatively ward. The basin floor in this instance is characterized by a
unconfined frontal splays (Posamentier and Kolla, 2003a) as minimal leveed channel and a relatively widespread frontal
illustrated in Figure 33. The available relief from point of flow splay or lobe. Where the flow has had a long run before reaching
origin to the basin floor can also play an important role in the basin floor (Fig. 34B) the flow is much better organized and the
determining the style of turbidite deposition. Assuming constant transition point lies farther across the basin floor.
slope gradient, the greater the relief from shelf edge to basin floor,
the longer the run of the turbidity flow, and therefore the greater DEPOSITIONAL ELEMENTS
the tendency for concentration of sand towards the flow base.
This allows greater efficiency of levee construction and a greater The integration of facies description and process sedimentol-
likelihood for levees to extend across the basin floor. All else ogy leads to the identification of larger-scale depositional or
being equal, two flows of identical grain-size composition flow- architectural elements and their linkage into depositional sys-
ing down two slopes with different length (each characterized by tems and ultimately depositional sequences. Depositional ele-
the same gradient), one with relief of a few hundred meters and ments in deep-water systems are of the order of ten to a few tens
the other with relief of a few thousand meters, can be associated of meters in thickness, and may extend laterally for tens of meters
DEEP-WATER TURBIDITES AND SUBMARINE FANS 421

Confined flow (within leveed channel) Frontal splay

Lower-density
part of flow
Levee top

Higher-density part of flow

Location of hydraulic jump


Abrupt flow disorganization
and expansion

FIG. 33.—Schematic depiction of flow expansion that occurs at the base of slope where the change of gradient is abrupt. Flow on the
slope is sufficiently well organized so that coarser-grained sediments are entrained in the flow base and are fully confined by
levees. Once the flow strikes the abrupt change in gradient that is located at the base of slope, the flow abruptly becomes
disorganized as it experiences a hydraulic jump. This abrupt increased disorganization causes a sudden increase in sand content
within the upper part of the flow and results in a situation where the higher-density part of the flow lies above the levee crests
and avulsions are likely to occur at this location. At this abrupt base-of-slope location, transition from leveed channel to frontal
splay is likely (compare with Fig. 32B, C).

Lower-density part of flow


A

Higher-density part of flow

Lower-density part of flow

Higher-density part of flow

FIG. 34.—Schematic depiction of two similar sediment gravity flows at the point of initiation (i.e., similar sand-to-mud ratio, volume,
etc.), but facing slopes of the same grade but significantly different length. The sediment gravity flow that faces the short slope
A) has less distance available to it for sorting to occur than does the sediment gravity flow facing a long slope B). The result is that,
when the flow finally reaches the basin floor, the flow down the long slope is significantly better sorted, with coarser sediment
more concentrated near the flow base, than the flow down the short slope, where poorer sorting results in sands much higher up
in the flow.
422 HENRY W. POSAMENTIER AND ROGER G. WALKER

up to tens of kilometers. The elements are defined by (1) their the canyon walls than are slope channels. In the distal reaches of
external geometries and (2) the internal facies within the ele- canyons, as channel-wall relief diminishes, turbidity current
ments. Our approach will follow that of Mutti and Normark height eventually exceeds the height of the canyon walls, and
(1991), and our analysis of depositional elements will integrate levees develop.
stratigraphic, geomorphologic, and facies observations, based on
seismic, outcrop, and borehole data. Moreover, the relationships Canyons.—
between depositional elements in time and space will be within
the framework of process sedimentology. We will start with An example of a canyon is shown in Figure 36. There is no
depositional elements that are observed in proximal settings— evidence of levee construction on the flanks of the canyons,
slope channels and canyons—and then move progressively far- suggesting that flows were fully confined within this feature. The
ther seaward down the slope and across the basin floor, where we presence of sand within this canyon is largely at the base (Figs. 37–
will examine leveed channels, overbank deposits (including sedi- 39), expressed as moderate- to high-sinuosity channel threads.
ment waves, crevasse splays, and planar levee deposits), frontal These channel deposits can be fully to partially preserved, the
splays (i.e., lobes or distributary channel complexes), and debris- latter illustrated by the segment of high-sinuosity channel depos-
flow deposits (including debris-flow lobes, channels, and sheets). its observed in the canyon terrace perched above the canyon floor
In each instance we will suggest, based on process sedimentol- (Fig. 37). The canyon fill is inferred to be overwhelmingly mud
ogy, what facies would be encountered in association with spe- dominated, as evidenced by the seismic reflection character
cific depositional elements. observed within the confines of the canyon Figs. 37, 40). The
seismic reflection pattern of canyon fill commonly is character-
Canyons and Slope Channels ized by moderate- to low-amplitude, discontinuous chaotic-con-
torted seismic reflections. This seismic reflection character com-
Canyons and slope channels are the primary conduits for monly has been associated with mass-transport deposits such as
sediments to travel from the shelf-edge staging area, across the associated with slides and debris flows (Posamentier and Kolla,
slope, and onto the basin floor. They can range in scale from a 2003a). The morphology of the top of the canyon fill is character-
few meters in depth and width (these would be referred to as ized by linear flow lines (Fig. 36A, C), further evidence for the
slope gullies) to ten or more kilometers wide and over a kilome- absence of turbulence in the flows responsible for at least the most
ter deep (submarine canyons). They tend to be largely erosional recent phase of canyon fill (Posamentier and Kolla, 2003a). How-
with significant incision into the substrate. In the context of this ever, there can exist isolated threads of channel sands, character-
discussion, the distinction we draw between canyons and slope ized by high-amplitude, continuous to discontinuous seismic
channels is that canyons consistently fully confine the flows that reflections at the canyon base or embedded within the canyon fill
pass through them, whereas slope channels only partially con- near the canyon base.
fine the flows that pass through them (Fig. 35). The effect of The fill of many canyons is commonly fine grained, and it is
partial confinement is that some spillover from the tops of the deposited after the canyon or channel has been abandoned. If
flows passing through slope channels occurs, resulting in the abandonment is due to cutoff of sediment supply during rela-
construction of levees on the flanks of the channel. Levee con- tive sea-level rise, the fill may consist largely of slump and slide
struction does not occur when the flows are fully confined, as material from the canyon or slope channel walls with additional
they are with canyons (though some canyons contain smaller contribution of hemipelagic mud and silt that gradually settles
leveed channels within the confines of the canyon walls). An- over the area of the slope and in the canyon. The present-day
other distinction between canyons and slope channels is that, Mississippi Canyon appears to have filled in this way (Goodwin
whereas both can be deepened by the passage of turbidity and Prior, 1989). Consequently, it is likely that the preponder-
currents, canyons are more likely to widen by mass wasting on ance of canyon filling occurs only after the axial turbidity-

A B

Canyon Slope Channel or Gully

FIG. 35.—Schematic illustration showing section views across a canyon A) and slope channel or gully B). Sediment gravity flows
within canyons are fully confined by canyon walls. Consequently no levee deposits are observed outside the canyon. In contrast,
sediment gravity flows within slope channels or gullies are not fully confined by channel walls so that levee deposits are observed
outside the channels or gullies.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 423

A B C

Flo
w
dir
ec
tio
n

Figure 42

10 km

FIG. 36.—Seismic attributes of the modern sea floor in the vicinity of Mississippi Canyon, Gulf of Mexico; seaward is from upper left to
lower right. A) seismic dip azimuth map—this map highlights the principal canyon as well as smaller tributary channels. The canyon
floor is characterized by long linear grooves. The walls of the canyon are characterized by delicate tributary networks of small gullies.
B) seismic dip magnitude—this attribute highlights the edges of the canyon floor as well as the edges of the tributary channels. C)
seismic reflection curvature—this attribute highlights the drainage networks and drainage divides; it also clearly shows the arcuate
nature of the canyon walls in places (compare with Fig. 42). Seismic data proprietary to PGS Marine Geophysical NSA.

Mass-transport deposits

Flow direction

FIG. 37.—Seismic cross section through Mississippi canyon, with low-sinuosity to moderate-sinuosity turbidite channels at the
canyon base. Note that the bulk of the canyon fill comprises mass-transport deposits. The axial channel shown here corresponds
to the channel segment shown in Figure 38A. The sinuous channel on the flank of the canyon fill is only partially preserved. Seismic
data proprietary to PGS Marine Geophysical NSA.
424 HENRY W. POSAMENTIER AND ROGER G. WALKER

A B
A B

1 km
C
C

1 km 1 km

Sinuosity = 1.2

Flo
w
Di
rec
tio Channel width = 200–250m
n

Sinuosity = 1.5
10 km

FIG. 38.—Seismic reflection amplitude images of the channel at the base of Mississippi canyon (compare with Figs. 37 and 39).
Sinuosity ranges from 1.2 proximally to 1.5 distally. Seismic data proprietary to PGS Marine Geophysical NSA.

FIG. 39.—Coherence image of proximal Mississippi canyon channel. Note the discrete channel threads indicate meander loop
expansion (i.e., swing) and down-system migration (i.e., sweep). Seismic data proprietary to PGS Marine Geophysical NSA.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 425

current channels shut down. Once these “conveyer belts” cease Some canyons can be completely mud filled and contain
to be active, any sloughing off the canyon walls remains within channels that have no associated sand deposits. This commonly
the confines of the canyon. Extensive slumping commonly occurs within tributary arms of larger canyon systems. Such
characterizes the canyon walls (Figs. 40, 41). Thus, the bulk of tributaries may have formed as a result of retrogressive slumping
canyon filling likely occurs as a result of mass wasting during on the canyon margin. The heads of these tributary systems do
rapid sea-level rises and subsequent highstands, thus making not apparently link up with shelfal fluvial systems; hence there is
the fill predominantly part of the transgressive and highstand no significant sand input. These tributary systems commonly are
systems tracts. characterized by small slope gullies or rills, feeding larger, com-
Some canyon systems characterized by supply of very coarse- monly straight axial channels. A tributary arm of Mississippi
grained sediment can have extensive coarse-grained sediments Canyon, shown in Figure 42, is characterized by channels at the
deposited within canyons (e.g., Pigeon Point Formation, Lowe, base but notably no associated high-amplitude seismic reflec-
1979; Carmelo Formation, Clifton, 1981, 1984). This is most com- tions, suggesting a complete absence of sand in the channels.
mon in active continental-margin settings where deep-water Such features are thought to have formed by low-density, slow-
turbidite systems are linked to short and steep fluvial drainage moving turbidity currents originating along the crests of steep
systems. drainage divides.
The facies that would most likely characterize the bulk of the
canyon fill is that of a debrite, that is, a mud-dominated deposit Slope Channels.—
with minimal internal organization. Convolute bedding associ-
ated with isolated cohesive blocks can be present in some Sand-prone slope channels such as that shown in Figure 43
instances. The isolated channel deposits (e.g., Fig. 37) observed have been described in some detail by Hackbarth et al. (1994),
within the canyon would be characterized by true turbidite Mayall and Stewart (2000), Kolla et al. (2001), and Posamentier
facies, likely dominated by Bouma A and B units, with Bouma (2003a). In contrast with the canyon previously discussed, this
C, D, and E units commonly lacking preservation potential type of sediment conduit is associated with levee construction.
because of erosion by successive turbidity currents through the Such levees can be observed high up on the slope at least to
channel. These channels commonly are not associated with within 8 km of the shelf-edge staging area (Posamentier, 2003a).
constructional levees, inasmuch as the canyon walls serve the The channel thread at the base of the slope channel illustrated by
purpose of confining the flows in their entirety, thus precluding Posamentier (2003a) is characterized by high-amplitude seismic
the possibility of flow spillover and levee construction. In reflections and a moderate- to high-sinuosity channel pattern
isolated instances, small channels within the middle to upper that persists landward nearly to the toe of slope of a small shelf-
part of the canyon fill can have associated levees, all deposited edge delta, which itself is laterally confined to the head of the
within the confines of the canyon walls. This situation can slope channel (Fig. 44). The presence of channel sinuosity nearly
develop when a turbidity flow travels across the relatively flat to the slope channel head suggests the presence of turbulent
floor of a partially filled canyon. Essentially, this constitutes an flow at least this high up on the slope, if not all the way to the
underfit situation; the flows coming through the channel no shelf–slope break. Further, a time slice through the upper slope
longer “feel” the canyon walls, hence they form their own and outer shelf reveals a protuberance of the shelf edge pre-
confining levees. cisely where the slope channel is located (Fig. 45). The associa-

A B

Note scallop-shaped
slump scars on
canyon wall

Canyon Margin
Flow direction

FIG. 40.—A) Transverse profile through Mississippi Canyon illustrating canyon cut as well as sand-prone channel fill at base of
canyon. Dotted line indicates location of time slice shown in B. B) Time slice through Mississippi Canyon illustrating arcuate walls
indicative of extensive slumping along canyon walls. Dotted line indicates location of time slice shown in A. Data proprietary to
PGS Marine Geophysical NSA.
426 HENRY W. POSAMENTIER AND ROGER G. WALKER

A B

on
cti
re
Di
Flo

w
w

Flo
Dir
ec
tio
n

5 km

FIG. 41.—A) Curvature map of modern sea floor over Mississippi Canyon characterized by arcuate walls indicative of extensive
slumping. B) Time slice across Mississippi Canyon and 3D perspective view of interpreted arcuate wall of the canyon. Data
proprietary to PGS Marine Geophysical NSA

5 km

FIG. 42.—Seismic reflection curvature of the sea floor along southwestern side of Mississippi Canyon, Gulf of Mexico (detail of Fig. 36).
Three larger channels up to one kilometer wide, as well as numerous smaller gullies and rills, are shown. Note that many of the
smaller gullies and rills originate along knife-edge drainage divides. Seismic data proprietary to PGS Marine Geophysical NSA.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 427

Flow direction

FIG. 43.—3D perspective illumination of the Einstein Channel on the upper slope of the Desoto Canyon area, eastern Gulf of Mexico.
Note the presence of levees external to this channel as well as the sinuous nature of the channel pattern. The channel is
approximately 1.5 km wide, from levee crest to levee crest. Seismic data courtesy of VeritasDGC.

tion of a shelf-edge protuberance or delta with the slope channel Sikkima and Wojcik (2000), and Abreu et al. (2003). The amalgam-
suggests a genetic link between shelf fluvial and distributary- ated sandy facies are characterized predominantly by Bouma A
channel sediment delivery systems and the presence of turbid- and B turbidites, whereas the underfit late-stage channels are
ite deposits on the slope. One can infer from this close connec- characterized by Bouma A to D turbidites. Climbing current
tion between the fluvial distributary channels and the meander- ripples can be observed most commonly within the levees asso-
ing threads within the slope channel that the turbidity currents ciated with the late-stage underfit channels.
which came through the slope channel may have originated as
density underflows (i.e., hyperpycnal flows) sourced by river- Examples of Coarse-Grained Canyon Fills.—
borne sediments. Such processes likely would have been active
for days or weeks at a time, while the river was in flood. Direct In unusual cases, the upper parts of large canyons may
links between fluvial distributary channels and canyon systems contain significant quantities of sand. Two examples have been
is also common. Most well-developed canyons are associated reported from the Atlantic margin of Brazil, in studies based on
with major river systems on the shelf (e.g., the Hudson canyon, closely spaced log and core control. The sand-prone fill of a large
the Congo canyon, and the Mississippi canyon) (Posamentier slope channel (canyon?) has been documented in the Carapeba–
and Allen, 1999). Pargo system on the Atlantic margin offshore Brazil (Fig. 46;
Slope channels commonly contain sand-prone facies at or Bruhn and Walker, 1995), where a large slope channel can be
near the base of the conduit. Mayall and Stewart (2000) document traced for at least 150 km. Turbidites have been studied in the
examples and propose a model for slope channel fill that contains Carapeba and Pargo oilfields, which lie about 90 km down-
debris-flow deposits at the base overlain by amalgamated turbid- canyon from the updip erosional edge of Cretaceous rocks. The
ites. The fill culminates with late-stage, isolated underfit leveed slope-channel width is of the order of a few kilometers, the
channels associated with diminishing flow discharge, located precise width of the channel being hard to determine because
within the master slope-channel walls. Other examples of com- younger turbidites spread more and more widely over the upper
plex slope channel fill are documented in Kolla et al. (2001), parts of the slope-channel margins. The thickness of the fill is
428 HENRY W. POSAMENTIER AND ROGER G. WALKER

Flow direction

Transverse
section
D

Axial
section

FIG. 44.—A) 3D perspective view of the Einstein channel (approximately 2 km wide) with associated sand-prone channel fill at the
base and shelf-edge delta at the slope-channel head. Each delta-front shingle is shown in a different color. B) Axial section through
the Einstein slope channel illustrating the shingled nature of the slope-channel-head delta as well as the high reflection amplitude
of both the delta as well as the channel fill at the base of the slope channel. C) Transverse section through the Einstein channel
illustrating high-amplitude sand-prone deposits at base of channel. D) Seismic reflection amplitude of the deposits at the base
of the slope channel illustrating the presence of sinuous narrow channel threads, characterized by down-system migration of
meander loops. Seismic data courtesy of VeritasDGC.

about 260 m (Turonian) plus 24 m (Maastrichtian). The turbidites hypothetical continuous outcrop of this canyon fill, 200 m thick
within the slope channel form eight thinning- and fining-upward and covering one km2, would be almost impossible to interpret as
successions. Individual beds at the bases of the successions can be a canyon fill—the only suggestion of channelization might be the
up to 12 m thick and consist of granule sandstones with scattered very thick beds and the coarse grain sizes.
pebbles (Fig. 47; the pebbly sandstone facies described earlier in The second example of canyon filling is also from the Atlantic
this paper). The successions are 27 to 140 m thick and can be margin of Brazil, in Regencia Canyon (Bruhn and Walker, 1997).
mapped as tabular or linguoid sandbodies 1 to 12 km wide in The canyon can be mapped near the mouth of the Doce River (Fig.
which the younger turbidites become finer grained, thinner 48) and is up to 6 km wide. It can be mapped for at least 15 km, and
bedded, and more discontinuous upsection and downcanyon. the fill is up to 1 km thick. Lagoa Parda field is about 2.8 km long
The successions are stacked in an overall retrogradational back- and 2.5 km wide, and it has 70 wells with an average spacing of 200–
stepping pattern for at least 20 km, recording the fill of the slope 250 m. Seven wells are cored, with a total of 324 m of core. Because
channel. of this unusually good control, individual beds can be traced from
The slope-channel morphology can be traced at least 60 km well to well; where beds or groups of beds can no longer be
downslope from Carapeba (Fig. 46), thus defining the deposi- correlated, channel margins can be defined (detailed correlations
tional site at Carapeba and Pargo as within the slope channel. The are shown by Bruhn and Walker, 1997, their fig. 13, and a schematic
depositional elements consist of the thinning- and fining-upward correlation diagram is shown here in Fig. 49). The thirty-eight
successions, which appear to spread with a sheet-like geometry channels so defined are 210 to more than 1050 m wide and over 1
from wall to wall (Bruhn and Walker, 1995, their Figs. 5 and 6). A km long. The channel fills range in thickness from 9 m to more than
DEEP-WATER TURBIDITES AND SUBMARINE FANS 429

e
ic
sl
e
m
Ti
Time slice

Shelf-Edge
Shelf-Edge Delta
Delta
B

FIG. 45.—A) Perspective view of Einstein channel and associated shelf edge, Gulf of Mexico, from 3D seismic data. The surface shown
lies at the base of the channel levees and is Late Pleistocene in age. B) Time slice through the upper part of the section reveals a
protuberance of the shelf edge corresponding to a shelf-edge delta (compare with Fig. 57). Note that this protuberance
corresponds precisely to the location of the Einstein slope channel. Seismic data courtesy of VeritasDGC.

50 m. Relationships of channels and their stacking patterns are channels generally being steeper than slopes facing away from
shown in Bruhn and Walker (1997, their figs. 14, 15, and 16). the channels. Levees are also steeper on the left sides of the levees
Channel fills comprise bouldery to pebbly conglomerates in (Coriolis effect in the southern hemisphere). In CC3, channel
normally graded beds up to 6.4 m thick, along with graded orientations again suggest flow from the northwest margin of
sandstone beds up to 3.8 m thick. Finer-grained facies include Regencia Canyon, and asymmetrical levee growth appears to
bioturbated mudstones and thin-bedded sandstones, and mo- have influenced channel switching (details in Bruhn and Walker,
notonous dark gray mudstones. Detailed correlations show that 1997). Channels are only 230–280 m wide, with fills 13–15 m thick.
these finer-grained facies are deposited as the levees of channels Two main points emerge from this study. First, it is clear that
filled with the coarser facies. turbidity currents can deposit coarse sediment within canyon
The channels can be grouped into three channel complexes heads, where the distance between flow initiation and sediment
(CC), colored orange (channels 1–11), yellow (13–35), and red deposition is only a few kilometers. Deposition may be strongly
(36–38) in Figure 49. Overall, the channel fills become narrower, influenced by the abrupt flattening of the gradient from canyon
thinner, and finer grained from CC 1 (orange) to CC 3 (red). margin to canyon floor, particularly in CC1 and CC3. Second, it
Channel complexes 1 and 2 are deeply incised. Channel orienta- is clear that small (10–20 m deep, 200–300 m wide) leveed chan-
tions in CC1 suggest flows from smaller tributaries along the nels can develop within canyon heads—this makes the interpre-
northwest margin of Regencia Canyon (Fig. 48), and there are no tation of some limited outcrops very difficult. Specifically, the
levee facies associated with the CC1 channels. Orientations in channels at San Clemente, California (described below), have
CC2 suggest flows generally from west to east along the axis of coarse fills, muddy channel walls, and multiple incisions. They
the canyon. Levees are associated with channels 19–35 (Fig. 48). have previously been described as mid-fan channels (Walker
Slopes on the levees are up to 10 degrees, with slopes facing the (1975b), but this interpretation is revised below, partly in the light
430 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 46.—Location map of Carapeba and Pargo fields, offshore Brazil. Note Brazilian coastline and city of Campos. Long serrated
line shows limit of occurrence of Upper Cretaceous rocks (after Bruhn and Walker, 1995).

of a resemblance between the San Clemente channels and the CC3 m thick, with indications that the nested channels may be as much
channels of Lagoa Parda. The general point to emphasize is that, as 40 m thick (Walker, 1975b). In addition to horizontal beds,
unless an ancient example can be fairly positively identified, it there are eight prominent inclined surfaces, seven of which are
cannot be used to construct general facies models—using the data draped with mudstone (Fig. 51). The surfaces have dips ranging
from San Clemente without being sure of its setting (canyon head from 5 to 18 degrees, and the mudstone drapes vary from 30 cm
or basin floor) could lead to distorted syntheses of data during to 2 m in thickness.
modern–ancient comparisons and facies modeling. The interpretation of the San Clemente section is that it
comprises a series of nested channels, separated by dipping mud-
Example of Slope-Channel Fill, draped surfaces. There were no outcrop indications that the
San Clemente, California.— channels might be sinuous, although excavations of channel
walls suggested that the strikes of the walls varied from 230 to 300
A candidate for a slope-channel fill is observed in the degrees (Walker, 1975b).
Capistrano Formation (Upper Miocene) at San Clemente, Califor- The fill of the channels consists of interbedded sandstones
nia (Fig. 50; Walker, 1975b; Campion et al., 2000; Camacho et al., and mudstones. The grain size is up to coarse sand, with many
2002). The channel complex cuts into mudstones containing some beds containing scattered granules and pebbles. Sandstone bed
beds of chert (similar to beds in the Monterey Chert). Bedding is thickness is variable; in some of the channels it is typically a few
generally horizontal in these mudstones, but some beds are tens of centimeters (Fig. 52), whereas in other channels beds can
broken by small soft-sediment faults suggesting movement on a be almost 1 m or thicker (Fig. 53). The thicker beds tend to be
slope—thus the beds outside the channel may be slope deposits. amalgamated, without mudstone layers (Fig. 54). Amalgamation
Within the channel complex, beds are very well exposed. surfaces are characterized by grain-size changes and bedding
They are also horizontal, and the channel complex can be traced irregularities that suggest loading of the upper sandstone into the
for over 500 m (Fig. 51). The Upper Miocene part of the cliff is 20 lower one (Fig. 54).
DEEP-WATER TURBIDITES AND SUBMARINE FANS 431

channel margin. In the case of the lower two beds, thin sandy
layers a few millimeters thick continue up the channel margin
and eventually disappear. It appears that the turbidity currents
transported sand close to the bed, and that silt and mud (with very
thin sand layers) was draped higher up over the channel margin.
Note also that potential reservoir rocks in the channel pinch out
laterally against the channel margin within a few meters.
The suggestion of soft-sediment disturbance in the beds out-
side the channel complex, and the proven occurrence of small
leveed channels within larger incisions (Lagoa Parda, Brazil;
Bruhn and Walker, 1997), suggests that the Capistrano Formation
at San Clemente can be interpreted as a slope-channel fill.

3D Seismic Examples.—

In contrast with outcrop data, good-quality 3D seismic data


can afford a comprehensive view of depositional elements and
their relationship to each other, though, of course, ground-truth
calibration may be lacking. In a well-documented example from
the eastern Gulf of Mexico, Posamentier (2003a) illustrates the
evolution of a linked shelf-edge delta and slope channel. The
shelf and slope environment became an active depocenter likely
in response to sea-level lowering and associated forced regres-
sion resulting in a seaward shift of the depocenter across the
shelf and to the shelf edge. Once the depocenter reached the
shelf edge a lowstand shelf-edge delta formed. This lowstand
delta, shown in axial view in Figure 57, is characterized first by
successive downstepping of the delta plain, likely associated
with falling sea level (i.e., the early lowstand systems tract of
Posamentier and Allen, 1999, or the falling stage systems tract of
Plint and Nummedal, 1998), and later aggradation of the delta
plain, likely associated with rising relative sea level (i.e., the late
lowstand systems tract of Posamentier and Allen, 1999, or
simply the lowstand systems tract of Plint and Nummedal,
1998). Numerous small slope channels or gullies are observed
on the surface that marks the base of this lowstand delta com-
plex (Posamentier, 2003a) (Fig. 58). These slope gullies are not
uniformly distributed along the entire breadth of the slope
10 cm shown, but rather tend to cluster in one area. The area where the
gullies are clustered directly coincides with the location of the
shelf-edge protuberance, suggesting a genetic link between
deltaic distributary channels and downdip slope gullies
(Posamentier, 2003a). These slope gullies are observed only at
the base of the lowstand delta complex; within and to the top of
the delta a single larger slope channel can be observed (Figs. 43,
44). This larger slope channel seems to have captured most of
FIG. 47.—Core photo of pebbly sandstone, Carapeba Field. Core
the flow from the shelf systems at the expense of the numerous
sleeves are 1 m long, top to left. The core shows one pebbly
smaller slope gullies that originally were present and seems to
sandstone bed, with prominent fluid-escape features in the
represent the culmination of an evolution from many slope
uppermost two core sleeves.
gullies to a single slope channel within a single depositional
sequence.
The finer layers between the sandstones consist of gray silt-
stones and mudstones with abundant bioturbation. On the in- Basin-Floor Leveed Channels
clined mudstone drapes, gray silty mudstones are irregularly
interbedded with very fine-grained brown claystones (Fig. 55). Basin-floor leveed channels commonly are genetically linked
The different fine-grained facies were not examined micro- with canyons or slope channels. Examples of basin-floor leveed
paleontologically, but commonly in Tertiary turbidites in Califor- channels are illustrated in Figures 59 and 60 and are described by
nia the gray silty mudstones contain transported shallow-water Posamentier et al. (2000), Peakall et al. (2000), and Posamentier
foraminifera whereas the brown claystones contain deeper-wa- (2003b). Aggradation of channel-fill deposits can occur both on
ter benthonic foraminifera. It therefore is assumed that the in- the basin floor (Fig. 61) and on the slope. The Joshua channel,
clined mudstone drapes consist of silt and mud introduced by described by Posamentier (2003b), built a channel ridge that
turbidity currents (the gray layers), as well as hemipelagic brown stands c. 65 m above the adjacent basin floor. Likewise, the
claystones deposited between turbidity currents. channel shown in Figure 61 stands c. 90 m above its adjacent basin
These relationships are particularly well displayed in channel floor. Sinuosity can be variable, from segments that are nearly
6 (Fig. 56). The lower three beds pinch out rapidly toward the straight to segments that can display sinuosity of up to 3.0.
432 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 48.—Location map of Regencia Canyon, State of Espirito Santo, Brazil. The canyon head is just onshore, close to the present day
Doce River mouth, 550 km northeast of Rio de Janeiro.

FIG. 49.—Diagram of channels within the Lagoa Parda field, Regencia canyon head. Orange—channel fills of unstratified bouldery
to pebbly conglomerate and very coarse-grained sandstone. Yellow and red—unstratified coarse-grained sandstone and parallel
stratified medium-grained sandstone. Green—interbedded bioturbated mudstones and thin-bedded sandstones, interpreted as
levee facies. Note decrease in channel size and grain size upward. Lower channels do not have levees, but channel horizons 26
and higher have associated levees.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 433

San Clemente
N
0 0.5 km

Capistrano Formation
San Clemente
Outcrop Belt State Beach

L o Afte
0 160 km

ca r W
(
tio a
n o lke
y
dar

f C r, 1
oun

h a 97
k B
Par

n n 5)
els
1–
San Francisco

8
Pa
Pa
cif

cif
ic

ic
Oc

Oc
ea

Los Angeles
n

ea
n
Study Area

FIG. 50.—Location of the outcrop at San Clemente, California, south of the parking lot (from Campion et al., 2000).

40

Bioturb. Mdstns. Bioturb. Mdstns. Pleistocene


Parking
20 Lot Cgls. WIDE GULLY Cgls. Terrace

Mainly massive ssts Mainly massive ssts U. Miocene


Capistrano Fm
0
0 Meters 50 100 150 200
8 7
40

Path from Campground


Bioturb. Mdstns. at Top of Valley

20 Cgls.
Cgls.

CU
Path
0
210 250 300 350 Restrooms 400
6 5 4 4 3 Path Tunnel
40 2
Lifeguard

Bioturb. Mdstns.

20 Cgls.
Lower cliff edge

Bioturb. Mdstns. Bioturb. Mdstns.

0 Bioturb. Mdstns.
450 Poorly exposed 500 550 600 650
Tunnel 2 1 Tunnel

Mudstone Drape

FIG. 51.—Measured section of the cliff at San Clemente. Distances are in meters south of the parking lot. Regional bedding is
horizontal, and heavy dipping black lines indicate mudstone drapes on channel walls (numbered 2 through 7). Channel 6 is shown
with red arrow and is discussed in the text. From Walker (1975b).
434 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 52 (above).—Four thin but coarse-grained sandstone layers


separated by gray (turbidite) mudstones. Part of the fill of
channel 7, San Clemente.

Meander-loop migration is a common attribute of deep-water


channels, with both meander swing and sweep being common
features (Fig. 62). This can be observed both in section as well as
in map view (Figs. 62–64). In many instances, the seismic expres-
sion of meander loops in plan view can be characterized by a
series of scrolls (Fig. 62). Somewhat less common are meander-
loop cutoffs with associated oxbows (Fig. 65). Internal scour at
successive channel bases within an overall aggradational amal-
gamated channel package is very common. The base of basin- 15 CM
floor channel complexes is erosional with at least some incision
into the substrate (Fig. 64).
In association with meander-loop channel migration, lateral-
accretion deposits can be observed on seismic sections (Figs. 66–
69). Abreu et al. (2003) present a well-illustrated example of lateral-
accretion sets in a deep-water meandering channel offshore west
Africa. Such features are similar in stratigraphic architecture to
fluvial point bars (Fig. 70), though the formation process can be
quite different. Nonetheless stratigraphic compartmentalization
from a petroleum flow-unit perspective is quite similar.
Lateral-accretion surfaces have been observed in outcrop as
well, as in the Ross Sandstone and overlying Gull Island Forma-
tion of western Ireland (Lien et al., 2003). Figure 71 shows a cliff
outcrop with prominent lateral-accretion surfaces characterized
by alternating layers of sandstone and mudstone. These lateral-
accretion surfaces have a vertical relief of about 5 m. In the Gull
Island Formation, lateral-accretion surfaces can be seen in a
continuously sandy succession (Fig. 72).
Lateral accretion commonly can be seen in 3-D seismic images
(Figs. 66–69). In Figure 62 the “point bars” show a series of scrolls


FIG. 53.—Graded coarse sandstone at San Clemente, part of the
fill of channel 8.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 435


FIG. 54.—Two coarse sandstone beds amalgamated along a
loaded contact (black line). Quarter for scale. Note gray
(turbidite) mudstones above and below.

that may be comparable to the mudstone-draped lateral-accre-


tion surfaces of the Ross Formation or to the channel-wall drapes
from San Clemente (Fig. 51).
In both the San Clemente and the Ross examples, the lateral-
accretion surfaces contain prominent mudstone layers. These
could form vertical and/or lateral baffles to fluid flow in a
reservoir situation. Meander-loop (“point bar”) settings in the
subsurface (Fig. 62) may therefore present engineering problems
with respect to fluid flow within the meander loops.
The process responsible for meander-loop migration may
involve a single flow event with lateral accretion developing in a
way similar to point-bar development in fluvial systems. That is,
a prolonged flow event would produce an undercut bank along
outer bends of a meandering channel, and at the same time
accretes sediment on inner bends, forming lateral-accretion sets
such as shown in Figures 71 and 72. An alternative explanation
for how deep-water meander loops migrate down system is
shown in Figure 73. This process involves a succession of discrete
flows, so that, with each flow event, cutting or erosion accompa-
nies the passage of the head and body of the turbidity flow, and
fill accompanies the tail. Each flow event would preferentially
erode outer bends more deeply, thus resulting in lateral shift of
channel axis (i.e., meander swing). Moreover, because the thal-
weg would exercise maximum erosive force on the outer bend
just down system of the channel bend itself, the meander loop

FIG. 55 (below).—Part of the drape on the wall of channel 2. The


gray layers are silty and easily disaggregated, and represent
mud introduced by the turbidity currents. The brown layers
are clay and contain a benthonic fauna. They represent hemi-
pelagic deposition between turbidity currents. Arrows indi-
cate truncation surface of the margin of channel 2 (see Fig. 51).
436 HENRY W. POSAMENTIER AND ROGER G. WALKER

REGIONAL BEDDING

RESERVOIR SECTION

FIG. 56.—Channel 6 at San Clemente, California. Note the lateral pinchout of the lower three beds against the channel margin (red
arrows). In detail, the beds can be traced part way up the wall, where they are represented by very thin (< 1 cm) sandy layers. The
yellow dotted lines mark the mudstone drape on the channel wall—note that the lower two sandstones pinch out into the drape.
In the subsurface, the thick turbidites in the channel fill might make up a reservoir section—but note how rapidly this section
would disappear along strike.

one km

100 msec

one km
FIG. 57.—Shelf-edge delta in close proxim-
100 msec ity to Einstein Channel. Note the suc-
cessive downsteps of the delta top, in-
Late lowstand dicating that the delta prograded un-
der the influence of falling sea level.
This is an excellent example of deposits
om plex associated with forced regression. The
sta nd c slightly progradational to aggrada-
e low tional section that caps the delta repre-
Bas sents deposition during the latter
phases of a sea-level lowstand, when
sea level is slowly rising. Seismic data
courtesy of VeritasDGC.
Early lowstand
(forced regression)
DEEP-WATER TURBIDITES AND SUBMARINE FANS 437

A C

Shelf Shelf

Slo Slo
pe pe

one km
B D (approximately)

FIG. 58.—A, C) Clustering of slope gullies on two late Pleistocene surfaces in the area of the Einstein Channel, Gulf of Mexico, derived
from 3D seismic data. In each instance the gully clustering is at the base of a lowstand slope succession, and in each instance the
top of the lowstand slope succession is characterized by a single, larger slope channel, not coincidentally where gully clustering
at the base was greatest. B, D) illustrate seismic reflection profiles parallel to and transverse to the slope gullies. The relief of these
gullies commonly is less than 20 m. Seismic data proprietary to PGS Marine Geophysical NSA.

Sea floor dips gently Sea floor dips steeply Sea floor dips gently
(< 1 degree) to the East (~ 6 degrees) to the East (< 1 degree) to the East

Flow Direction

2 km
N
FIG. 59.—Seismic reflection dip-azimuth map of deep-water leveed channel system in the Makassar Strait, Indonesia. The sea floor
dips to the right at a very low angle (less than 1 degree) with the exception of a seaward-dipping segment as indicated. This steep
dip is associated with the surface expression of a syndepositional base-of-slope toe thrust. Note the abrupt increase in channel
sinuosity and decrease of channel width seaward of this toe thrust. Note also the extensive sediment waves on both sides of the
leveed channel (compare with Fig. 77) (after Posamentier et al., 2000). Seismic data courtesy of WesternGeco.
438 HENRY W. POSAMENTIER AND ROGER G. WALKER

Flow
Dire
ction

5 km

FIG. 60.—Seismic reflection dip-magnitude map of Joshua channel. The channel itself is part of a channel belt that lies within a larger
leveed channel (shown by red arrows). (Posamentier, 2003b). Seismic data courtesy of WesternGeco.

Levees

Flow direction

Crevasse splay 4 km
(approximately)
Leveed channel

FIG. 61.—3D view of the top of a Pliocene leveed channel with crosscutting seismic section.
Note the crevasse splay deposited upon the lower overbank. Seismic data courtesy of
WesternGeco.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 439

Flo
w
dir
ect
ion
Sweep

3
2
1
Swing

FIG. 62.—Seismic horizon slices illustrating meander-loop


swing and sweep within the Joshua Channel, Gulf of
Mexico. Channel threads indicate order of formation one km
(channel 1 is first, channel 3 is last). Seismic data courtesy
of WesternGeco.

n
r e ctio
w di
Flo

Sweep
1

3 one
one km
km
4

5 Amplitude extraction - horizon slice (-36 msec)

FIG. 63.—Seismic horizon slice through the deep-water leveed-channel system shown in Figure 59, in the Makassar Strait, Indonesia
section. Down-system migration of meander loop is well expressed. Channel threads indicate order of formation (channel 1 first,
channel 5 last) illustrating channel sweep. Seismic data courtesy of WesternGeco.
Passive channel fill 440
Channel fill
A
A Flow direction
C

Levee
Levee
A'

Frontal splay deposits

A' Channel
Channel base
base A

one km

D
HENRY W. POSAMENTIER AND ROGER G. WALKER

No vertical exaggeration

one km
FIG. 64.—Seismic reflection profiles across the channel–levee system shown in Figures 59 and 63 (illustrated in plan view in A) and section view in B– D). This section
shows the channel fill to be characterized by high-amplitude reflections, inferred to be sand-prone (C). The transverse profile also is shown without vertical
exaggeration to reinforce the observation that interpretation of the architecture of these types of deposits can be greatly facilitated by “squashing” the profile. Seismic
data courtesy of WesternGeco.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 441

A B
Incipient neck cutoffs

Flo
w
di r
ec
tio
n
D
five km

C D

one km one km

FIG. 65.—Meander-loop development within the Joshua channel, eastern Gulf of Mexico, as observed on seismic reflection horizon
slices (compare with Fig. 62). A) Time slice that illustrates two incipient cutoffs of meander-loop necks. B) Slightly higher
(approximately 12 m) time slice that illustrates cutoff of meander-loop necks. C, D) Time-slice details of the neck cutoffs and
resulting oxbows shown in Parts A and B. Seismic data courtesy of WesternGeco.

would therefore display down-system migration (i.e., meander Channel-fill lithofacies in these basin-plain leveed channels
sweep). Figure 73 illustrates how flow velocity vectors tend to be would be similar to that which is encountered within slope chan-
greater along outer bends especially just down-system from each nels. Those basin-floor channels characterized by more significant
meander bend, thus tending to cause a down-system shift of aggradation would allow greater preservation of waning-phase
successive channel axes (i.e., meander sweep). The resulting turbidites, so that Bouma A to C units would be most common.
architecture of cut and fill can readily appear as lateral-accretion
architecture on seismic data, whereas the outcrop expression Levee and Overbank Deposits
would be one of cut and fill accompanying a progressive shift of
channel axes. Within each cut and fill, it is possible that smaller- Channel levees commonly are deposits with concave-up,
scale true lateral accretion such as shown in Figures 71 and 72 can gull-winged shapes. In general, the facies common to this depo-
be present. sitional element include CCC turbidites, lenticular bedding and
Avulsion events in deep-water channels have been docu- small-scale erosion, slumping involving one or two beds, and
mented using 3D seismic data (Posamentier and Kolla, 2003b). large-scale chaotic failures. Seismically, levees tend to be charac-
Figure 74B, C illustrates a late lowstand leveed channel that terized by low-amplitude continuous to discontinuous reflec-
underwent an avulsion event. Avulsion events tend to be asso- tions (Figs. 64, 67, 69, 74).
ciated most commonly with levee breaches or crevasses on Because the tops of turbidity currents associated with channels
outer channel bends (Figs. 61, 75), though in isolated instances, commonly ride higher than the channel walls, there is continual
because of flow perturbations, crevasses can form on inner spillover onto surrounding areas (Fig. 27). When this occurs, flows
bends as well. abruptly become less confined, and in response to this flow expan-
442 HENRY W. POSAMENTIER AND ROGER G. WALKER

A
Flo
w
Dir
ec
tio
n

one
one km
km

100 ms

one km

Lateral accretion

Slice level

FIG. 66.—Seismic reflection horizon slice illustrating A) plan view and B) section view across a high-sinuosity deep-water leveed
channel, Gulf of Mexico. Down-system migration of meander loops (A) and lateral-accretion deposits (B) are clearly shown.
Seismic data courtesy of WesternGeco.

sion (i.e., increase in cross-sectional area), flow velocity abruptly tion, and ripped-up mud clasts (CCC turbidites). These sedimen-
decreases. Lowered flow velocity results in decreased sediment- tary structures are more common in proximal overbank/levee
carrying capacity; that is, the flow is less capable to transport settings than in any other turbidite environment of deposition.
sediment, and rapid sedimentation out of suspension takes place. Enhanced spillover occurs at outer bends of channels as the
As noted before, thin-bedded turbidites in levees commonly con- upper parts of flows tend to continue in a straight-line trajectory
tain climbing current ripples (Bouma C facies), convolute lamina- whereas the lower parts of flows tend to follow the curving
DEEP-WATER TURBIDITES AND SUBMARINE FANS 443

A x y
FIG. 67.—Seismic reflection horizon slice illustrat-
ing A) plan view and B–C) section view across
the Joshua Channel shown in Figure 60. The
section views (B and C) clearly illustrate mean-
der-loop migration and lateral accretion. The
lateral accretion in this system is characterized
by a strong aggradational component. Seismic
X' data courtesy of WesternGeco.
ction
one
one km
km Flow Dire Y'

B C
x X' y Y'

Slice level

100 msec one km

FIG. 68.—A) Seismic reflection horizon slice


and B, C) section views across a leveed
A channel system, Gulf of Mexico. Both the
horizontal as well as the vertical sections
clearly show down-system meander-loop
X'
X' migration as well as lateral accretion (see
yy yellow boxes and line-drawing interpreta-
tions). The lateral accretion in this system is
characterized by minimal aggradation. Seis-
mic data courtesy of WesternGeco.

x Lateral accretion x’

n
ctio
w dire
Flo
x
x y y’

Y'
Y'
one
one km
km

B x
x X'
X'

Slice level

100 msec
C yy Y'
Y'
Slice level

one km
444 HENRY W. POSAMENTIER AND ROGER G. WALKER

A 55 km x y FIG. 69.—Seismic reflection horizon


km
slice through a meandering slope
channel, A) offshore Nigeria, il-
lustrating down-system migration
of meander loops. The two seis-
mic reflection cross sections B, C)
illustrate the lateral-accretion
tio
n packages that are associated with
irec meander-loop evolution. Seismic
d
w data courtesy of VeritasDGC.
Flo
x' y'

x' x
B
Lateral accretion

Slice level

y' y

C
Lateral accretion

100 msec 5 km

A B
Flo
n wd
tio ire
irec ctio
n
d
ow
Fl

C
nn
iioo
cctt
irree
i
dd
ooww
FFll F IG . 70.—Fluvial examples from
Wyoming showing plan-view
expression of lateral accretion,
meander-loop migration, and
meander-loop cutoffs. Channel
widths are approximately 500 m.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 445

LATERAL ACCRETION

BEDDING

FIG. 71.—The Ross Sandstone in Rehy Cliff, Shannon Estuary, western Ireland. Regional bedding here is horizontal, and the dips in
the sandstone and mudstone layers in the upper part of the cliff are interpreted to represent lateral accretion on a “point bar”-
like feature in a submarine channel. Vertical relief of the lateral accretion surfaces is about 5 m.

B C

2:1 Vertical exaggeration


FIG. 72.—A) Lateral-accretion surfaces in a sandstone from the Gull Island Formation (immediately above the Ross Formation) of
western Ireland. Thickness of sand body about 20 m. B) Two-to-one vertical exaggeration of outcrop photo with line-drawing
interpretation C).
446 HENRY W. POSAMENTIER AND ROGER G. WALKER

A
B

Flow direction
Time 1

Time 2
Slice Level
Time 3

Time 1 Time 2 Time 3

C
D
Flow-velocity contours Turbidite deposits
Time 1 Time 1 Cut – Initial phase
Fill – Late phase

Time 2 Time 2

Time 3 Time 3

FIG. 73.—Moderately sinuous deep-water channel, eastern Gulf of Mexico. Cross section view shows apparent lateral accretion A).
Seismic time slice shows meander loop expansion (i.e., swing) as well as down-system meander-loop migration (i.e., sweep)
B). Schematic cross section illustrates flow velocity contours with highest velocity shown in yellow. Note the progressive
lateral shift of channel axis through time in response to highest flow velocity occurring on the right side of each channel C).
Schematic depiction of successive turbidite channel fills. Note that with the progressive lateral shift of the channel axis through
time, only the left side of each channel (with the exception of the final channel position) is preserved D). Seismic data courtesy
of WesternGeco.

channel path. This process of enhanced spillover by which the Outcrop Example—Simi Hills, California.—
upper part of the flow tends to shear off and decouple from the
lower part of the flow has been referred to as flow stripping (Piper An excellent example of levee deposits in outcrop is the
and Normark, 1983). One result of flow stripping is that levee Cretaceous Chatsworth Formation, Simi Hills, California, which
crests on outer channel bends are consistently higher than those has been described by Link et al. (1984). Several hundred meters
on inner channel bends (Fig. 76). Flow-stripping deposits in of the Chatsworth Formation can be viewed along the north side
certain instances can take the form of transverse sediment waves of the freeway (Highway 118) that traverses the Santa Susana
(Fig. 77–80). These sediment waves consistently appear to be Pass (Fig. 82). Thick-bedded, channelized turbidites can be exam-
thicker on their proximal flanks, suggesting up-system migration ined in Chatsworth State Park, with channel–levee facies exposed
of wave crests (Fig. 78). The sedimentology of some sediment just to the south.
waves has been documented by Migeon et al. (2000) in the Var The levee facies is characterized by interbedded sand-
fan-channel turbidite system (Fig. 79). These waves also are stones and mudstones in beds up to a few tens of centimeters
characterized by up-system wave-crest migration and have formed thick. Sandstone grain size is up to medium sand. Beds in the
transverse to flow direction. Cores taken from the updip, crestal, basal parts of thinning-upward successions tend to be struc-
and downdip limb of one of the Var sediment waves reveal tureless and amalgamated (Fig. 83), but the thinner beds at the
significantly greater amounts of sand present in the updip limb of tops of successions are commonly lenticular (Fig. 84) and are
the sediment wave and the least amount in the downdip limb characterized by CCC turbidites (Fig. 8). Climbing ripples, and
(Fig. 79). Grain sizes up to medium sand were observed in the climbing ripples that become convoluted, are well displayed
updip limb. Other sediment waves observed in mid- to upper- in Figure 7. A scour with ripped-up mudstone clasts is shown
slope settings, which are not associated with turbidity-flow chan- in Figure 85. The combination of these features suggests an
nels but rather with oceanic currents such as loop currents, likely environment in which minor erosion is common, yet indi-
are not sand prone (Fig. 81). vidual beds are fairly coarse though relatively thin. Also, the
DEEP-WATER TURBIDITES AND SUBMARINE FANS 447

Avulsion node
A B

Flo
wd
irec
tion

G
C D
Leveed channel

Slices
A–F

Frontal splay
Condensed section

100 msec one km


E F

FIG. 74.—Succession of horizon slices through a deep-water turbidite system, eastern Gulf of Mexico. A–F) Horizon slices illustrate
the evolution of this system from an initial frontal splay (F), eventually evolving into an isolated leveed channel (A). The cross-
section view (G) illustrates this evolution as well. Seismic data courtesy of WesternGeco.

climbing ripples and convolute lamination suggest rapid depo- combine to characterize the upper parts of channel margins, or
sition from suspension. The combination of features described the backs of levees. These two settings can best be distinguished
suggests rapid deposition of relatively coarse sand, yet in a by their context, as shown in the next example.
setting in which the beds are consistently relatively thin—that
is, high on a channel margin, or on the back side of a levee. In Outcrop Example—Wheeler Gorge, California.—
this setting, turbidity currents may in places be erosive (Fig.
85), and the bedding lenticularity may be a function of depo- An outcrop example of distal thin-bedded turbidites overlain
sition in minor erosional scours and hollows (Fig. 84), or due by a channel–levee complex is observed in the Cretaceous Wheeler
to the pinching of beds against a depositional topography (the Gorge channel systems north of Ojai, California (Rust, 1966;
channel margin or levee). Walker, 1975b). The beds are vertical and are preserved as a sliver
Evidence for local slope instability within the levee deposits along the Santa Ynez Fault of the Transverse Ranges. They occur
is in the form of slump deposits involving a few beds, with as an isolated outcrop, and cannot be related to other Cretaceous
undeformed beds above and below. Such slumps are commonly deposits in southern California. This example emphasizes the
associated with CCC turbidites, and in this context probably importance of the vertical relationships of depositional elements
represent the local sliding of one or two beds, either on the back in establishing an overall interpretation of the depositional set-
of the levee away from the channel or down the channel wall ting. Once the overall setting has been established, individual
toward the channel. In many instances these slumps are sub- depositional elements can be reexamined in the light of the
seismic in scale. However, a seismic-scale example of slumps overall setting.
into a channel is illustrated in Figures 76 and 80, and slumps on The section can be subdivided into three parts (Walker, 1975b):
the distal side of a levee in Figure 86. (1) a lower section of “zebra-striped” mudstones with 1-cm-scale
The important architectural elements in these settings consist graded siltstones (about 250 m thick; Fig. 87); (2) a central section
of CCC turbidites, lenticular bedding styles, and slumps involv- that consists of three conglomerate-to-sandstone packages (105
ing just one or two beds. Singly or in combination, these elements m thick; Fig. 88); and (3) an upper section of interbedded sand-
B 448

A A'
one km
Note positive relief

C
Channel fill eroded
50 m B B'

A Avulsion channel
Avulsion node

Note positive relief

A
A Note negative relief

B
B
Note negative relief
HENRY W. POSAMENTIER AND ROGER G. WALKER

B'
B'

A'
A'
Knickpoints
Flow direction

FIG. 75.—A) Shaded oblique perspective view of channel shown in Figure 76. Two knickpoints are shown in association with an avulsion channel. The channel down-
system of the kickpoints remains unfilled and is characterized by a concave-up profile C). Upstream of the knickpoints the channel is filled with sand as indicated by
the convex-up profile B). Seismic data courtesy of WesternGeco.
Relief: ~ 6–7 m 625 m
Avulsion node 2
A A' A
B Fl
ow
di
re
ct
io
n

100 msec

Avulsion node 1

Avulsion node 2
A' D
A

Slump scars
DEEP-WATER TURBIDITES AND SUBMARINE FANS

five km

FIG. 76.—Images of Pleistocene Joshua channel, eastern Gulf of Mexico (Posamentier, 2003b). The perspective view B) and detail C) reveal the elevated nature of the channel
fill, suggesting a differential compaction effect associated with predominant sand fill of the channel. Note that levees are highest on outer bends. Note also the small
slump scars that characterize the inner levee slope. The channel is part of a channel belt that has accreted to a level 65 m above the adjacent basin plain D). Internally,
the channel is characterized by lateral accretion coupled with vertical aggradation A). Seismic data courtesy of WesternGeco.
449
A 450

Flow direction
one km

B Sediment waves

C
HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 77.—Leveed channel on the basin floor of the Makassar Strait, Indonesia C), characterized by overbank sediment waves illustrated on the associated dip magnitude
map A). Sediment waves are best developed on outer bends of the channel. Compare with the schematic illustration B). Seismic data courtesy of WesternGeco.
A Direction of wave migration
A'

100 m Accretion

Current direction

one km

A A'
DEEP-WATER TURBIDITES AND SUBMARINE FANS

FIG. 78.—Sediment waves associated with slope channels offshore Nigeria. Note that the levee crests migrate progressively up-system, suggesting landward migration
of these sediment waves. Seismic data courtesy of VeritasDGC.
451
Overspill direction 452

A B
E7° 45 E7° 55
NW KN127 KN126 KN124 SE
Depth
(meters)

2400
N43° 30

Wave migration direction

2500
Overspill

KN127
KN126
KN124

2400
N43° 20

2500
1 km

Profile NIC34

KN127 KN126 KN124


HENRY W. POSAMENTIER AND ROGER G. WALKER

1m
F IG . 79.—Overbank sediment
waves from the Var system,
Mediterranean Sea, shown in
plan view A), section view B),
and core C) (from Migeon et al.,
2000). Note the up-system wave
migration (B) as well as the pres-
ence of sand that characterizes
primarily the upstream limb of
these waves.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 453

Sediment waves

Flow direction

Slump scars

one km

FIG. 80.—Sediment waves in the overbank of the Joshua channel, eastern Gulf of Mexico, shown on a curvature map extracted from
the upper bounding surface of the channel–levee complex. Note also the small-scale slump scars on the inner levee. Seismic data
courtesy of WesternGeco.
454 HENRY W. POSAMENTIER AND ROGER G. WALKER

A B

Direction of wave migration


Shelf/slope break
Shelf/slope break
D
C

one km

one km

FIG. 81.—Sediment waves on the middle to upper slope of the eastern Gulf of Mexico. These waves appear to be migrating obliquely
upslope. They are not associated with any nearby channel or canyon. Wavelength ranges from 400 to 600 m and wave amplitude
ranges from 5 to 10 m. Seismic data courtesy of VeritasDGC.

FIG. 82.—Outcrop of Cretaceous Chatsworth Sandstone on north side of Highway 118, Santa Susana Pass, California. Note subtle
changes in bed thickness and sand/shale ratio from highway to skyline.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 455

FIG. 83.—Sharp-based (yellow arrow), thick-bedded sandstones, with one prominent amalgamation surface (red arrow). The
irregular horizontal marks were made by an excavator! Total thickness of structureless sandstone about 3 m. This photograph
was taken about 20 years ago at the back of a trailer park—it is unlikely that the outcrop still exists; if it does it is probably
overgrown. Chatsworth Sandstone near Chatsworth, California.

FIG. 84.—Interbedded sandstones and mudstones in the Chatsworth Sandstone, stratigraphically a few meters above the structureless
sandstones of Figure 83. Note the lenticularity of many of the beds (arrows). Thickness of section about 5 m.
456 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 85 (above).—Scour outlined in yellow, with coarser-grained


structureless sandstone fill and abundant ripped-up mud-
A stones clasts. Chatsworth Sandstone near Chatsworth, Cali-
fornia.

FIG. 86.—Slump scar on levee of deep-water channel–levee


complex in Makassar Strait, Indonesia. Seismic data courtesy
of WesternGeco.

C
B
Flow direction

Slump scar

one km
DEEP-WATER TURBIDITES AND SUBMARINE FANS 457

Cgl. Seq. 1
Mudstones
Zebra Mudstones FIG. 87.—Pace-and-compass map of
the lower part of the Wheeler
Gorge section on Highway 33,
north of Ojai, California. Note that
bedding is essentially vertical, and
that by rotating the map it also
Mudstones, 90 m serves as a measured section. The
section is slightly disrupted by
faulting, but the thicknesses shown
probably are fairly close.

Zebra
Mudstones, 120 m

Slump

50 m
North
Santa Ynez Fault

FIG. 88.—Pace-and-compass map of the central channel complex at Wheeler Gorge, California. The channel complex can be divided
into three separate conglomerate sequences, interpreted as one channel fill.
458 HENRY W. POSAMENTIER AND ROGER G. WALKER

stones and mudstones with abundant soft-sediment deformation As mentioned above, context is one of the most important
(more than 120 m thick; Fig. 89). characteristics that allow architectural elements to be assigned to
The zebra-striped mudstones consist almost entirely of color- depositional environments. The interbedded sandstones and
graded siltstones 1 to 2 cm thick alternating with mudstones (Fig. mudstones directly overlie the graded coarser facies of the chan-
90). Right at the base, these beds are slumped in a section some 5 nel fill. If the channel had migrated laterally, shifting the main
meters thick. The entire package is faulted, but it is estimated that coarse-sand depocenter elsewhere, the channel margin could
there are about 5500 of the graded siltstone beds. None of the migrate over the channel fill. A seismic example of this relation-
characteristics of levee deposits is present (i.e., no climbing ship is shown in Figure 64B.
current ripples, no mud rip-up clasts, etc.); the depositional The third conglomerate-to-sandstone package is about 30 m
environment is postulated to be very distal—either at a fan fringe thick, and consists of conglomerates (25 m) that grade into
or lateral but very distal to a channel. structureless sandstones (5 m) (Fig. 97). These are in turn overlain
The central section comprises three stacked conglomerate-to- by thin-bedded classical turbidites (Fig. 97). The base of this
sandstone packages. The base of the first sharply overlies the package is spectacular, and erodes about 5 m into the underlying
zebra-striped mudstones and is characterized by large flute casts beds. Large stratified sandstone–mudstone clasts have been ripped
(Fig. 91) and associated groove casts, some of which retain from the channel wall and deposited in the channel, with a sand
pebbles at the ends of the grooves. There are also abundant and conglomerate matrix (Fig. 98). Conversely, coarse sand from
ripped-up clasts of the zebra mudstones within the basal channel the channel has been injected into the beds of the channel wall as
fill (Fig. 92). This lowest package is about 25 m thick, and it sills (Fig. 99)—the sill in this photograph terminates abruptly, but
consists of conglomerates and interbedded structureless sand- smaller sills splay from the top and bottom of the main sill. In
stones and is interpreted as channel fill. The conglomerates are places, the sills feed dikes, clearly showing that these coarse
characterized by abundant mudstone clasts, probably derived deposits are not original beds in the channel wall. Well developed
from adjacent channel walls. The second package is approxi- sills and dikes have also been documented in the subsurface (the
mately 50 m thick. There is a basal conglomerate, but the bulk of Tertiary Alba Field in the North Sea; Hurst et al., 2005) as well as
the succession consists of graded conglomerate–sandstone beds in outcrop (Surlyk and Noe-Nygaard, 2003; Hurst et al., 2005).
(Fig. 93) and graded pebbly sandstones (Fig. 94). The uppermost The uppermost part of the section exposed in Wheeler Gorge
part of this package overlies these channel-fill deposits and consists of about 120 m of interbedded sandstones and mud-
comprises a 15-m-thick succession of thin-bedded turbidites and stones. The lower 80 m consists of classical turbidites, with beds
mudstones. Convolute lamination (Fig. 95), bedding lenticularity averaging about 10 cm in thickness. Overall, there is a thinning-
(Fig. 96), and minor erosion surfaces (Fig. 96) with small ripped- upward succession. Individual beds have small ripped-up mud-
up mudstone clasts suggest that this is a channel-margin or back- stone clasts, and convolute lamination is common. There is no
of-levee facies. soft-sediment deformation. The upper 40 m contains similar

FIG. 89.—Pace-and-compass map of the section above the channel complex at Wheeler Gorge, California. Two separate levee
complexes have been identified (both within the slumped turbidites).
DEEP-WATER TURBIDITES AND SUBMARINE FANS 459

TOP

FIG. 90 (above).—Very thin-bedded “zebra-striped” mudstones


below the main channel complex (Fig. 87) at Wheeler Gorge,
California. The striped appearance is the result of the stacking
of a large number of thin graded siltstone (pale) to mudstone
(dark) beds. Flat lamination and ripple cross lamination are
rare.

classical turbidites, but soft-sediment deformation is abundant.


A few beds a meter or so in thickness are completely disrupted,
but more commonly the deformation is restricted to a few beds
that show distinct soft-sediment folds (Figs. 100, 101). Convolute
lamination is also present, and in places there are distinct trends
in bed thickness—two thinning-upward successions, one thick-
ening-upward succession, and one thickening-to-thinning up-
ward succession.
The sedimentary folds involving one or two beds, and the
presence of convolute lamination, suggest that these thin-bed-
ded turbidites are channel-margin or levee deposits. Unlike the
channel-margin deposits described above, these thin-bedded
turbidites are not closely associated with thick, coarse-grained
channelized deposits. They are interpreted to represent back-of-
levee deposits associated with a channel system different from
that described above—a channel that is not exposed at Wheeler
Gorge (Fig. 102).


FIG. 91.—Large flute casts on the base of the lowest conglomerate
sequence (Fig. 87), Wheeler Gorge, California. The bed is
vertical, flow direction is shown by the yellow arrow, and the
highway tunnel gives the scale.
460 HENRY W. POSAMENTIER AND ROGER G. WALKER

MUDSTONE
CLASTS

BASE OF CONGLOMERATE

TOP

BASEMENT CLASTS

FIG. 92.—Very sharp base of conglomerate sequence 1 on the underlying zebra-striped mudstones. Note pale-colored basement clasts
and abundant ripped-up mudstone clasts.

TOP

FIG. 93.—Graded conglomerate-to-sandstone bed, as part of the fill of conglomerate sequence 2 (Fig. 88), Wheeler Gorge, California.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 461


TOP
TOP
FIG. 94.—Graded pebbly sandstone, part of conglomerate se-
quence 2 (Fig. 88). Irregular base shown by yellow dotted line.
Wheeler Gorge, California.

Outcrop Example of Levee Failure—New Zealand.—


Probably the best example of levee failure and collapse is in
the Waitemata Group (Eocene, New Zealand; Ballance, 1964).
The large-scale soft-sediment deformations have been described
by Gregory (1966). The best outcrop is on Whangaparaoa Head,
where there is about 4 km of continuous coastal exposure in high
cliffs. The facies consists dominantly of thin-bedded classical
turbidites, with soft-sediment folding occurring on many dif-
ferent scales. Some of the deformation consists of stacked dis-
harmonic folds that can be seen in the cliff (Fig. 103) and on the
wave-cut platform (Fig. 104). In other places, entire sections of
turbidites appear to have been rotated, as in Figure 105, where
the regional dip of the Waitemata Group is only 20 degrees. The
beds in this rotated section contain CCC turbidites (Fig. 106) and
slumps involving one or two beds (Figs. 25, 107); both of these
features strongly suggest a levee origin for the thin-bedded
turbidites, and hence a levee-failure origin for the disharmonic
folding and rotation. Locally, sections of thick-bedded turbid-
ites suggest associated channel fills (Figs. 108, 109).

FIG. 95 (below).—Convolute lamination in the uppermost part of


the fill of conglomerate sequence 2 (Fig. 88), Wheeler Gorge,
California.

TOP
462 HENRY W. POSAMENTIER AND ROGER G. WALKER

TOP

FIG. 96.—Bedding lenticularity (red arrow) in the interbedded sandstones and mudstones, uppermost part of conglomerate sequence
2 (Fig. 88). Yellow dotted line shows scouring and more bed lenticularity. Wheeler Gorge, California.

CONGLOMERATE, SEQUENCE 3

STRUCTURELESS SANDSTONES

THIN-BEDDED TURBIDITES

FIG. 97.—Overview of conglomerate sequence 3 (Fig. 88) after a forest fire. Note overall change from conglomerate, via structureless
sandstone, into classical turbidites. Section is about 30 m thick. Wheeler Gorge, California.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 463

LEVEE, SEQUENCE 2
TOP

FIG. 98 (above).—Base of conglomerate sequence 3 (Fig. 88),


shown by yellow dotted line. Conglomerate cuts into levee
deposits of sequence 2 (yellow arrow). Channel fill consists of TOP
basement clasts, as well as stratified clasts plucked from the
channel wall (pink arrows). Green arrow shows a sill of coarse
sandstone injected into the fine-grained levee deposits.
Wheeler Gorge, California.

Crevasse Splays
Another form of overbank deposit is the crevasse splay. In
contrast with sediment waves, which are associated with flow
over the levee crest and onto the overbank, the crevasse splay is
associated with flow through the levee and into the overbank
environment. Because the crevasse splay involves flow through
the levee, this flow taps deeper into the main flow than simple
spillover flow, which taps only into the upper part of the flow. As
a result, flow through a breach in the levee is sourced by the more
sand-prone part of the main flow.
A typical crevasse splay is characterized by a short channel
leading away from the main channel and feeding a smaller
distributary channel system (Figs. 61, 110–112). Note the plan-
view similarities between deep-water turbiditic and shallow-
water, continental-shelf deltaic crevasse splays (Fig. 113). A cre-
vasse splay can be considered a failed avulsion channel. The
distinction is that, in the case of an avulsion channel, flow is
permanently diverted through the crevasse and associated chan-


FIG. 99.—Sill of coarse sand and granules (see Fig. 98) injected into
interbedded sandstones and mudstones of the wall of channel
3. Note very abrupt termination of the sill. Wheeler Gorge,
California.
464 HENRY W. POSAMENTIER AND ROGER G. WALKER

REGIONALTOP
CONVOLUTE LAMINATION

FIG. 100.—Distinct soft-sediment folds involving two beds, with undisturbed bedding above and below. Note convolute
lamination in the uppermost bed of this small thickening-upward succession. Levee facies (Fig. 89) above the channel complex
in Wheeler Gorge.

REGIONAL TOP

FIG. 101.—Re-folded fold involving one bed in an otherwise well-stratified succession. Regional top shown by yellow arrow, and
way-up of the folded bed is shown by yellow arrows. Levee facies (Fig. 89), Wheeler Gorge, California.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 465

Wheeler Gorge

ps
rbank slum
Levee-ove
retion
ateral acc
L

Channel-Levee
System 2
retion
ateral acc
L
Vertical stacking

retion Channel-Levee
al acc
Later System 1

Basin floor zebra-striped mudstones

Fact Interpretation

FIG. 102.—Interpretation of the Wheeler Gorge section. The vertical line (“fact”) shows the location of the Wheeler Gorge section—
the rest of the diagram is a reconstruction based on Wheeler Gorge data plus models derived from channel shingling patterns in
fans such as the Amazon and the Rhone.

FIG. 103.—Large scale, disharmonically stacked soft-sediment folds in the Eocene Waitemata Group, Whangaparaoa Head, New
Zealand. Cliff is about 30 m high.
466 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 104.—View from cliff down onto the wave-cut platform, showing disharmonic soft-sediment folding. Width of section at
base of photo about 60 m.

REGIONAL DIP = 20o


70 m

FIG. 105.—The Eocene Waitemata Group has a regional dip of about 20 degrees—this photograph shows a section some 70 m thick
rotated by slumping to an angle of about 70 degrees. It lies just north of the disharmonic folds of Figure 101. The nature of the thin-
bedded turbidites in this outcrop is shown in Figures 106 and 107.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 467

CLIMBING
RIPPLES

TOP

CONVOLUTE
LAMINATION

FIG. 106.—Thin-bedded turbidites with convolute lamination and climbing ripples, exposed on the wave-cut platform as part of
the 70 m thick rotated section shown in Figure 105. Whangaparaoa Head, New Zealand.

SLUMPED BEDS

FIG. 107.—Thin-bedded turbidites, with one deformed horizon involving about 2 thin beds in a soft-sediment fold. Same location
as Figure 105.
468 HENRY W. POSAMENTIER AND ROGER G. WALKER


FIG. 108.—Thick-bedded turbidites in the Eocene Waitemata
Group at Whangaparaoa Head, New Zealand. 15 cm scale is
circled.

nel, whereas with the crevasse splay the flow diversion is tempo-
rary and relatively short-lived. In some instances, as revealed by
3D seismic data, the early stage of a crevasse splay is character-
ized by a levee breach that feeds a field of transverse sediment
waves (Fig. 114). As the system becomes progressively better
organized, the sediment waves are overlain by a gradually ex-
panding distributary channel network (Posamentier and Kolla,
2003b; Van Wagoner et al., 2003). Channels within this distribu-
tary network are associated with low-relief, probably sand-prone
levees.
The most distinctive sedimentary structure associated with
crevasse-splay deposits is climbing current ripples. As with
overbank sediment waves, flow expansion, which occurs when
the confined flow within the main channel cuts through the levee
and becomes unconfined within the overbank environment, re-
sults in rapid sedimentation from suspension. In addition, be-
cause of erosion through the levee, locally derived mud rip-up
clasts can be common. The stratigraphic architecture of crevasse-
splay deposits is characterized by amalgamated turbidites near
the apex of the splay, becoming less amalgamated with distance
away from the splay apex.
In general, crevasse splays are far more common in basin-
floor environments than in slope environments. On the slope,
gravity-flow vectors tend to be directed parallel to flow, down the

FIG. 109 (below).—Thick-bedded turbidites, with a suggestion of


a thinning-upward succession, in the Eocene Waitemata
Group, Whangaparaoa Head, New Zealand.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 469

n
tio
ec
dir
ow
Fl
Five km

FIG. 110.—Seismic reflection horizon slice illustrating a crevasse splay associated with a channel–levee complex on the basin floor of
the Gulf of Mexico. Note that several crevasses seem to have formed along this outer channel bend, feeding multiple small
channels characterized by bifurcation. Seismic data courtesy of WesternGeco.

one km
FIG. 111.—Seismic reflection horizon slice illustrating a crevasse splay associated with a channel–levee complex on the basin floor of
the Gulf of Mexico. Seismic data courtesy of WesternGeco.
A 470

one
one km
km
C
Leveed
Leveed Channel
Channel
Line
Line D
D

Crevasse Splay

Line
Line E
E
one km

Leveed
Leveed Channel
Channel
HENRY W. POSAMENTIER AND ROGER G. WALKER

Crevasse
Crevasse splay
splay

Crevasse
Crevasse splay
splay
E

FIG. 112.—Crevasse splay associated with a basin-floor channel/levee complex, Gulf of Mexico. Seismic reflection horizon slices A, B) illustrate deeper and shallower
slices through this system respectively. The crevasse splay is illustrated in perspective view in Part C) (see Fig. 61). Cross-section views through the principal
channel D) and the crevasse splay E) illustrate the high-amplitude character of these depositional elements. The high-amplitude character suggests the presence
of sand in these deposits. Seismic data courtesy of WesternGeco.
A
B

one km

C
DEEP-WATER TURBIDITES AND SUBMARINE FANS

one km
one km

FIG. 113.—Comparison between A) a shallow-water deltaic crevasse splay (from the Mississippi River delta) and B, C) two deep-water crevasse splays (Figs. 110 and
114) suggesting a strikingly similar morphology and likely similar depositional processes. Seismic data courtesy of WesternGeco.
471
A B C 472

Flow direction

Sediment waves

D E
HENRY W. POSAMENTIER AND ROGER G. WALKER

one km

FIG. 114.—Seismic reflection horizon slices from base A to top E) showing the morphological evolution of a leveed-channel crevasse splay. Note that the early
morphology appears to be characterized by transverse sediment waves as indicated by arrow (A). As the system evolves it appears to become better organized,
developing a distributary pattern (B through E). Seismic data courtesy of WesternGeco.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 473

relatively steep slope (i.e., slope angle commonly from 2 to 4 As discussed above, frontal splays form at the transition
degrees), rather than normal to flow parallel to contour, thereby point, the location where levee height is no longer capable of
diminishing the likelihood that sufficient force is exerted on fully confining the sand-prone or high-concentration part of the
channel walls (i.e., levees) to form crevasses and hence crevasse turbidity flow. At that location the sand-prone part of the flow
splays. Consequently the crevasse-splay sand habitat is an un- readily flows over the levee and establishes new courses on the
common element in slope settings. associated overbank. Also at this location, the down-system
trend towards increasing sand-to-mud ratio in the flow changes
Frontal Splays (i.e., Lobes) to a decreasing sand-to-mud ratio, as shown schematically in
Figure 26. Up-system of the transition point mud is preferen-
Leveed channels commonly are associated with terminal tially lost from the flow by overbank spillover, whereas down-
lobes, or frontal splays (Figs. 115–120). The terms frontal splay, system of the transition point sand is lost more rapidly from the
distributary channel complex, and lobe or lobeform can be used flow by sedimentation in the overbank and within distributary
somewhat interchangeably. Each term describes a depositional channels.
element that lies at the end of a leveed channel and tends to be fan Frontal splays have much in common with crevasse splays
or lobe shaped in plan view. Each term implies a different both sedimentologically and architecturally. Both are distin-
perspective on this feature: (1) frontal splay has process signifi- guished by relatively common current ripples and sheet-like
cance, wherein a flow spreads out or expands, (2) distributary bedding geometry with shallow channels and sand-prone, low-
channel complex has map-pattern significance, wherein succes- relief levees. Both tend to be more amalgamated near the splay
sive channel bifurcation results in a distributive channel network, apex and less so distally. Preservation of interbedded mudstones
and (3) lobe or lobeform has morphologic significance wherein becomes progressively more common with increasing distance
the deposit has a fan-shaped or lobate planform. away from the transition point, which corresponds to the frontal

Frontal splay

Transition point

Flow direction

five km

FIG. 115.—Frontal splay at the end of a leveed channel system on the basin floor of the Makassar Strait, Indonesia. The transition point
between leveed channel and frontal splay is located where levee height has diminished to below seismic resolution (compare with
Fig. 28). Seismic data courtesy of WesternGeco.
474 HENRY W. POSAMENTIER AND ROGER G. WALKER

Y Y'

A
100 msec
Slice level
Frontal splay
one km

one km Y'

FIG. 116.—Frontal splay at base of channel–levee complex, Makassar Strait, Indonesia. This splay complex is characterized in section
view by continuous to discontinuous high-amplitude seismic reflections and in plan view by extensive bifurcation. Note that, as
with the channel-levee complex illustrated in Figure 117, the frontal splay immediately underlies a moderate- to high-sinuosity
solitary leveed channel. Seismic data courtesy of WesternGeco.

splay apex. However, several characteristics distinguish frontal The channels appear to be multiple (possibly “braided”), al-
splays from crevasse splays: (1) Frontal splays tend to be larger, though it is not clear if more than one channel is occupied and
insofar as they involve the entire flow discharge in contrast with active during any single turbidity current. Other images show
only part of the flow discharge (i.e., that part of the flow that is that in more distal settings, the top of the frontal splay may be
temporarily diverted through the crevasse) associated with cre- essentially smooth—any topography is too small to be imaged.
vasse splay. (2) Crevasse splays commonly tend to be associated
with levee-derived rip-up clasts. These clasts tend to be mud Outcrop Example—County Clare, Ireland.—
prone. (3) Crevasse splays lie in close proximity to channel levees,
where the flows have passed through a breach. (4) Slope instabil- The outcrop example of a frontal splay discussed here is from
ity with resulting bed convolution and slumping is more com- the Carboniferous Ross Sandstone of western Ireland (Figs. 122,
mon in crevasse-splay settings, insofar as they are deposited on 123; Martinsen et al., 2000; Lien et al., 2003). The Ross is about 460
potentially steeper slopes of the overbank environment. m thick, and details of the regional geology, stratigraphy, and
Most commonly, frontal splays are deposited on basin floors paleotectonic setting are given by Lien et al. (2003). It can be
or on the floors of intraslope basins, with leveed channels and informally divided into lower and upper parts, 170 m and 290 m
channel complexes being the more common form of turbidite thick, respectively.
element encountered in slope environments. In some instances, Frontal splays are best seen in the lower Ross—the base is
however, frontal splays can be observed on slopes as well, espe- gradational from the underlying Clare Shale. Initial turbidites are
cially in intraslope basins (Prather et al., 1998), but also in some very thin and widely spaced stratigraphically, but overall, through-
instances on open slopes as well (Fig. 121). out the 170 m thickness, the beds tend to thicken upward (Lien et
The splays shown in this paper are mostly channelized, with al., 2003). In detail, however, there are no systematic thickening-
channel widths from about 100 m to smaller than can be imaged. upward or thinning-upward trends (on the scale of ten meters, or
DEEP-WATER TURBIDITES AND SUBMARINE FANS 475

Slice level

Transition point

FIG. 117.—Cross section A) and horizon slice B) views of a deep-water leveed channel feeding a frontal splay, Gulf of Mexico. Note
the clear location of the transition point where the leveed channel transitions into a frontal splay. The frontal splay is characterized
by continuous to discontinuous high-amplitude seismic reflections in section view, and by extensive bifurcation and braiding (?),
forming a complex distributive network in plan view. Seismic data courtesy of WesternGeco.

a few tens of meters), and no readily recognizable channels. Beds spillover lobes developed during the lateral migration of chan-
appear to be continuous for at least 200 m and of constant nels. The top surfaces of many of the packages show giant
thickness as far as they can be traced in the cliffs (Fig. 124). These erosional features resembling flutes (Fig. 127). The packages are
characteristics suggest that deposition was from unconfined superficially very similar to the thickening-upward sequences
turbidity currents that could spread on the sea floor with no first described by Mutti and colleagues (Mutti and Ricci Lucchi,
apparent topographic obstructions—characteristic of the distal 1972) and interpreted to result from the progradation of deposi-
parts of frontal splays. The absence of bed-thickness trends tional lobes. However, in the case of the Ross thickening-upward
suggests that each turbidity current was unrelated to the previ- packages, the intimate relationship with channel fills suggests
ous ones, with the deposits of larger and smaller flows interbed- that they are spillover (sediment waves?) rather than frontal
ded. The overall upward bed thickening suggests seaward step- lobes.
ping of the frontal splay over a long period of time. The channels in the upper Ross vary from about 10 to over 25
The upper Ross is characterized by thickening-upward pack- m and consist dominantly of amalgamated thick-bedded turbid-
ages, channel fills, slump–slide horizons, and turbidites with no ites. No consistent patterns of bed-thickness change were ob-
bed-thickness trends. Lien et al. (2003) showed that the thicken- served. Two areas define the relationships of channels to pack-
ing-upward packages (Figs. 125, 126) represented aggrading ages best—Kilbaha Bay and Cloonconeen.
476 HENRY W. POSAMENTIER AND ROGER G. WALKER

Frontal
Frontal splay
splay

Flow direction

Transition point

Leveed channel
five km

FIG. 118.—Horizon slice view of a deep-water frontal splay, Gulf of Mexico. The transition point is located where the levees associated
with the up-system leveed channel have diminished in height to below seismic resolution. The frontal splay is characterized by
extensive bifurcation, forming a complex distributive network. Seismic data courtesy of VeritasDGC.

At Kilbaha Bay (Fig. 123) almost horizontally dipping beds the packages outside the channel. This relationship strongly
are exposed for 1.5 km along the cliff, with a composite strati- suggests that the thickening-upward packages represent aggra-
graphic thickness of about 40 m. There are 20 thickening-upward dation of spillover lobes rather than frontal splays developed
packages (Fig. 128). In several locations the thick-bedded por- downstream from channel mouths. Both the Kilbaha and the
tions of the packages can be traced laterally and seen to split into Cloonconeen channels suggest that the thickening-upward pack-
non-amalgamated, thin-bedded turbidites with mudstone part- ages are closely related to channel filling. Lien et al. (2003) also
ings between beds. The Kilbaha Bay channel (Fig. 128) has a illustrate lateral-accretion surfaces, suggesting that the channels
visible depth of incision of about 3 m, with a fill of at least 6 m. The can migrate laterally (Figs. 71–73).
base is characterized by a layer of mudstone clasts, with at least These relationships between architectural elements suggest
two stratified sandstone blocks up to 3.5 m long and 35 cm thick; the following interpretation. Where the channel is far from a
these are interpreted to represent channel-wall collapse, as de- particular depositional site, the only overbank deposits to reach
scribed above from Wheeler Gorge. Most importantly, on the that site consist of mudstones (Fig. 131). During lateral channel
eastern side of the channel fill, thick-bedded amalgamated chan- movement toward the site, overbank deposition may consist of
nel-fill sandstones split into thinner beds and grade laterally into thin-bedded turbidites. The closer the channel approaches, the
adjacent packages (Fig. 128). This relationship suggests that the thicker and more amalgamated the overbank succession become.
packages are in some way related to channel filling (and not lobe Closest to the channel, turbidity currents may be scouring (form-
progradation)—the relationships are seen even better at ing the giant flutes) and bypassing without depositing.
Cloonconeen Point. The detailed relationships between all of the architectural
The channel at Cloonconeen Point (Figs. 129, 130) has a elements are shown in Figure 131. In phase 1 the channel is active
minimum fill thickness of 15 m. There appear to be at least three and is migrating laterally, forming lateral-accretions deposits on
separate incisions, and the fill consists of separate thick-bedded one side and eroding a cut bank on the other side. In phase 2
turbidites close to the margin but an almost completely amalgam- during channel filling, note the lateral shift of the cut bank,
ated succession closer to the channel center. The uppermost part resulting in thicker and thicker overbank deposits.
of the fill consists of about 3 m of amalgamated sandstones, but
these beds can be walked out laterally for about 100 m (Fig. 130), Integrated Interpretation of the Ross Sandstone
where they split progressively and pass into thin-bedded turbid-
ites separated by mudstone partings. These thin-bedded turbid- Lien et al. (2003) suggested that the upper Ross consists
ites are organized into a thickening-upward package identical to mainly of sinuous channels stacked into sinuous channel belts.
North South
Levee crests
A

100
Slice level msec

Levee Levee
one km

Channel

North South
B Frontal splay margin
Frontal splay margin

Slice level

100
msec
A
A B
B
DEEP-WATER TURBIDITES AND SUBMARINE FANS

one km

one km

FIG. 119.—Seismic reflection cross-section views through the frontal splay B) and the feeder leveed channel A) associated with the system shown in Figure 118. Note the
single leveed channel clearly shown in section view (A), contrasted with the high-amplitude continuous to discontinuous seismic reflection character of the frontal
splay. The section view through the frontal splay also shows that this system is characterized by numerous small channels that seem to coalesce into a sheet-like
morphology. Seismic data courtesy of VeritasDGC.
477
478 HENRY W. POSAMENTIER AND ROGER G. WALKER

Transition point

FIG. 120.—Leveed channel feeding frontal
splay in the Bering Sea offshore Alaska
(from Kenyon and Millington, 1995).

Flow direction

5 km

Frontal splay

FIG. 121 (below).—Leveed channel feeding frontal splay on the


mid-slope, eastern Gulf of Mexico. This 3D perspective
view illustrates a reflection amplitude map draped onto
the upper bounding surface of the channel–levee complex.
Note that in the final stages of this channel-frontal splay
Leveed channel complex a solitary leveed channel flows across the top of
the frontal splay. Seismic data courtesy of VeritasDGC.

Flow direction

Transition point

Frontal
Frontal splay
splay
DEEP-WATER TURBIDITES AND SUBMARINE FANS 479

FIG. 122.—Location map for the Ross Sandstone of western Ireland. Rectangle shows location of detailed map in Figure 123 (after
Lien et al., 2003).

FIG. 123.—Detailed location map for the Ross Sandstone (yellow) and the overlying Gull Island Formation (brown), western
Ireland (see Fig. 122) (after Lien et al., 2003).
480 HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 124.—Classical turbidites in the Lower Ross Sandstone. Note the bedding continuity and absence of any thickening- or
thinning-upward successions. Cliff north of Ballybunion, about 10 m high.

FIG. 125.—Thickening-upward packages shown by yellow arrows in the upper Ross Sandstone, south side of Ross Bay. Note
succession from mudstones into thin-bedded turbidites, with thicker-bedded amalgamated turbidites in the upper parts of the
packages.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 481

Scoured surface

FIG. 126.—Thickening-upward packages at Kilbaha Bay. Note the mudstones at the base, overlain by thin-bedded turbidites, overlain
in turn by amalgamated thick-bedded turbidites. The uppermost surface is scoured, but the scours are not visible in this picture.

11 m
m

FLOW DIRECTION

FIG. 127.—Giant flute at Ross Bay. The shape of the flute is shown by the yellow arrows, and the flow direction by the red arrow.
The flute scours into amalgamated sandstones with a top surface covered in sinuous-crested ripples. The flute itself is partly
filled with mudstones and thin-bedded turbidites, with one surface showing straighter-crested ripples. There are no ripples
on the steeply dipping walls of the flute. Note the scale—the flute is nearly 2 m wide.
482 HENRY W. POSAMENTIER AND ROGER G. WALKER

LATERAL PASSAGE FROM THICK-BEDDED TO THIN-BEDDED TURBIDITES

10 m

CHANNEL GIANT SCOURS

FIG. 128.—Cliff face at Kilbaha Bay. Beds can be walked out along the cliff and in the wave-cut platform. Mudstones and thin-bedded
turbidites are yellow, and amalgamated sandstones orange. Note lateral passage from amalgamated turbidites into thin-bedded
turbidites in several locations (blue arrows). Note also the giant scours (black arrows). The thick-bedded amalgamated turbidites
that fill the channel also pass laterally into thinner beds.

THICKENING-UPWARD PACKAGE LATERAL TO CHANNEL

THICKENING-UPWARD PACKAGES THICK-BEDDED SANDSTONES –


CUT BY CHANNEL CHANNEL FILL

FIG. 129.—Diagram of channel at Cloonconeen Point. Exposure is almost 100 percent, but details of the channel wall are partly
obscured by rock rubble and minor tectonics. Channel is filled with amalgamated structureless sandstones and cuts into adjacent
thickening-upward packages. Note that the uppermost amalgamated sandstones of the channel fill can be walked out laterally
into an overbank thickening-upward package.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 483

SANDIER-UPWARD PACKAGE

CHANNEL FILL

FIG. 130.—Channel margin at Cloonconeen point. Photographer is standing in the middle of the channel fill (two yellow arrows). The
channel margin is outlined in yellow (see Fig. 129), and the lateral passage from thick amalgamated channel sandstones to a
thickening-upward package (red arrow) is shown with a dotted yellow arrow. Person for scale at end of dotted yellow arrow.

PHASE 1
Lateral-accretion distance Lateral-accretion distance

Lateral-accretion distance


Thin-bedded turbidites

Vertical
aggradation
mostly mudstone

Overbank spill during lateral Lateral migration with minimum Sandstones and mudstones on gently
migration of channels channel-floor aggradation dipping lateral accretion surface

Amalgamated beds
PHASE 2 Thick- to thin-bedded turbidites
Thin-bedded turbidites
Giant flutes within amalgamated sandstones Thick-bedded fill on Surface of erosional
Mudstones
and on erosional bypass surface cut-bank side of channel bypass with giant flutes Onlap of vertical
accretion deposits
Mudstone drape on
Ross Bay bypass surface

Clooconeen Rehy Cliff


Complete
sandier-upward package Rinevella Lateral-accretion deposits
Sandstones pass laterally into mudstones
as they onlap the lateral-accretion surface

FIG. 131.—Composite diagram for the upper Ross Sandstone, showing the development of a channel. In phase 1, note the lateral
shifting, with lateral accretion on one side, and a cut bank on the opposite side. Thickening-upward sequences form as the
channel migrates, with mudstones deposited when the channel is far away (blue dotted line), thin-bedded turbidites as the
channel approaches (red dotted line), and thick-bedded amalgamated turbidites when the channel is closest (green dotted
line). As the channel migrates, the opposite bank may receive lateral accretion deposits. In phase 2, the main channel is filled
with structureless amalgamated turbidites, probably resembling those of Figure 68.
484 HENRY W. POSAMENTIER AND ROGER G. WALKER

The thickening-upward packages represented overbank spill, splay, with the non-packaged parts of the upper Ross being
and the non-packaged turbidites represented deposition on a deposited on smoother parts of the splay.
smooth basin floor far from any channels. The exact position of
the channels and channel belts was not discussed. The model is Importance of the Ross Formation Case History.—
shown in Figure 132, where three individual sinuous channels
are shown diagrammatically within a sinuous channel belt. At The importance of this case history is that it illustrates how a
the end of the sinuous channel belt, turbidity currents spread stratigraphic unit can first be subdivided into architectural ele-
laterally to form the unchannelized lower Ross turbidites. ments—non-packaged turbidites, thickening-upward packages,
In the light of the seismic images presented here, particularly channel fills, and slump–slide depositional units. The channels
those that show channelization in the proximal parts of frontal are the easiest individual element to interpret, and the interpre-
splays, it is possible that the entire Ross Sandstone represents a tation of the thickening-upward packages follows from their
frontal splay. The lower Ross would represent the smooth outer detailed relationships to the channel fills. Reliance on earlier
portion of the splay, with the overall thickening-upward succes- models that would suggest prograding depositional lobes would
sion resulting from gradual progradation of the splay. The upper give an alternative (and probably incorrect) interpretation. The
Ross would then represent the proximal channelized part of the suggestion that the upper Ross might represent the proximal part

Area of non-packaged
and poorly packaged
tabular turbidites

Area of
non-packaged and
poorly packaged
tabular turbidites

Sinuous
Sinuous belt
belt
reestablished
reestablished after
after
sea-level
sea-level fluctuation
fluctuation Sinuous belt width ~ 5 km
(marine
(marine band)
band)
Sinuous belt of
channels and packages

FIG. 132.—Overall interpretation of the Ross Sandstone. Individual sinuous channels in the upper Ross (shown as red, green and
blue) are stacked into sinuous channel belts (yellow). At curves in the individual channel, overbank spill can create thickening-
upward packages (red, green and blue “lobes”). Thus the sinuous channel belt consists of two main depositional elements—
channel fills and overbank spills. At the downstream end, the channels lose their topography and feed a smooth basin plain
(the lower Ross). The passage from lower to upper Ross implies progradation of the sinuous channel belts over the smooth
basin plain.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 485

of a frontal splay is based on stratigraphic context and compari- of salt domes and mud volcanoes also have been observed, and
son with 3-D seismic images. Thus an interpretation has been on the flanks of turbidity-flow channel levees (Figs. 80, 86).
constructed that relies on defining architectural elements, defin- Because in most instances mass transport originates in low-
ing their particular stratigraphic relationships, and comparing energy environments, the associated deposits of such processes
this construction with seismic images of the deep sea floor. In this tend to be mud-rich.
way, we build our own model or interpretation without reliance In plan view these deposits can be lobate or channelized (Figs.
on a preexisting model—the theme developed by Walker (this 145–148). In some instances mass-transport deposits can oppor-
volume) in the introduction to this volume. tunistically use the channel of an earlier-formed turbidity-flow
channel (Fig. 149). The seismic character in both planar horizontal
Debris-Flow Deposits—Mass-Transport Complexes or vertical section commonly is chaotic to contorted (Figs. 135,
150–153), though in some instances they can be characterized by
Various forms of debris-flow and mass-transport deposits large-scale convolute bedding (Fig. 154).
comprise common depositional elements in many deep-sea envi- Where debris flows encounter obstructions or where flow
ronments (Fig. 133). These deposits can assume a variety of velocities diminish abruptly, as is common near their termini,
shapes and sizes. A characteristic that seems common to all is internal compressional structural features (i.e., low-angle thrust
their highly erosional nature and their contorted, chaotic, and faults) can be common (Figs. 151,155, 156). On the basin floor, a
structureless internal architecture. Erosional relief at the bases of commonly observed aspect of mass-transport deposits is an
such deposits can exceed 250 m in some instances (Figs. 134, 135). apparent large-scale channelization that is not so much related to
Where clasts were embedded in the flow base, erosion of the erosion of a channel and subsequent fill as it is a consequence of
substrate is marked by long, linear striations or grooves that tend “plowing” of the sea floor by a plug of debris analogous to the
to diverge in the downsystem direction (Figs. 21, 136–140). Clast effect of a shovel pushing through a snow layer (Figs. 157, 158). In
sizes can range from cobbles to clasts larger than houses (Fig. 141). instances such as these it is possible to calculate the travel distance
In some instances, small outcrop-scale scour can also be observed of the mass-transported debris (Fig. 159). Lateral compression
beneath mass-transport deposits (Fig. 142). caused by multiple phases of mass-transport events also has been
Large mass-transport deposits have been observed to origi- observed (Figs. 152, 160).
nate at shelf edge (Fig. 143) as well as mid-slope locations (Fig. The hallmark of the lithofacies of mass-transport deposits is
144), where fault scarps as high as 54 m mark the point of slope its lack of organization. Commonly they comprise mud-sup-
failure. Smaller mass-transport deposits originating on the flanks ported conglomerates, though they can also occur as pure mud-
stone containing muddy rip-up clasts. In isolated instances they
can also be relatively sand-rich (Jennette et al., 2000).
The upper bounding surfaces of mass-transport complexes
can vary from smooth to highly rugose. In some instances this
surface rugosity can exert an influence over subsequent turbidity
-flow deposits (Fig. 161), whereas in other instances, because the
short wavelength of the bathymetric lows, only ponding of
mudstones seems to occur (Fig. 162).

SEQUENCE STRATIGRAPHY

A typical deep-water depositional sequence on the basin floor


100 msec has been proposed by Posamentier and Kolla (2003a) as consist-
ing of basal debris-flow material, overlain by sand-rich frontal-
splay deposits, in turn overlain by isolated leveed-channel de-
posits and finally by debris-flow deposits and a condensed
section. This sequence is in part a distillation of the work of
Weimer (1991), Piper et al. (1997), Pirmez et al. (1997), Manley and
Flood (1998), Maslin et al. (1998), Beauboeuf and Friedmann
(2000), Brami et al. (2000), and Winker and Booth (2000) (Fig. 163).
It is relatively unlikely that each of these stratigraphic units
would be observed at any given location; debris-flow deposits are
most common on basin floors and within canyons, frontal splays
are most common on basin floors, and leveed channels are com-
mon in both slope and basin-floor environments. In contrast,
condensed sections are widespread and ubiquitous. Figure 74G
shows a section of a deep-water depositional sequence, character-
one km ized by frontal-splay deposits overlying a condensed section (thin
stratigraphic unit corresponding to the transgressive and high-
stand systems tract of the preceding sequence). In this example,
MTD frontal-splay deposits gradually give way to an isolated leveed
channel. We suggest that this transition from frontal splay to
FIG. 133.—Seismic section from the eastern Gulf of Mexico illus- leveed channel during the waning phase of a deep-water deposi-
trating the extensive nature of mass-transport deposits. At tional sequence occurs because of a progressive decrease in sand-
this location, the mass-transport deposits constitute approxi- to-mud ratio within the flows, which originate at the shelf-edge
mately 45% of the total section shown. Seismic data courtesy staging area. This diminished overall sand content results in a
of WesternGeco. seaward shift of the transition point as discussed above.
A 486
Line C B

Line D

Flow direction one km

five km

South North
C

Slice level

100 msec
five km
West East
D
HENRY W. POSAMENTIER AND ROGER G. WALKER

Flow direction
Slice level

Terminal wall

FIG. 134.—A) Mass-transport complex in the basin floor environment, eastern Gulf of Mexico, which lies within a broad, flat-floored channel C) that ends abruptly
at a terminal wall D). These sediments likely did not travel far but rather comprise a mass of material that was pushed from behind and slid to some degree
along a decollement surface at the base (note the relatively flat base in section view (C and D). The complex rheology of this deposit is illustrated by the chaotic
nature of the seismic reflections within the mass-transport complex as well as the tongues of sediments that extend beyond the terminal wall B). The relief of
the channel is approximately 240 m. Seismic data courtesy of WesternGeco.
12.5 km

240 m

Flo
wd
irec
tion

30 km
DEEP-WATER TURBIDITES AND SUBMARINE FANS

Terminal wall

FIG. 135.—Perspective-view image of the base of the mass-transport complex shown in Figure 134. Seismic data courtesy of WesternGeco.
487
488 HENRY W. POSAMENTIER AND ROGER G. WALKER

Flow
Flow direction
direction


FIG. 136.—Grooves beneath mass-transport complex deposited
in a continental-slope environment (image courtesy of D.
Mosher).

FIG. 137 (below).—Grooves beneath mass-transport deposit at


Five km the base Paleocene, North Sea, in a basin-floor environment
(after Wilson et al., 2005).

Flow
Flow direction
direction

10 km
DEEP-WATER TURBIDITES AND SUBMARINE FANS 489

one km

Grooves

FIG. 138.—Dip-azimuth map of surface at the base of a channelized mass-transport deposit. Note the grooves that characterize this
surface and the tendency for these grooves to diverge down-system. This suggests that internal deformation characteristic of
flow rather than slide processes have occurred. Seismic data courtesy of WesternGeco.

Basal grooves

Outrunner blocks

one km
FIG. 139.—Grooves at the base of a mass-transport complex, basin floor, eastern Gulf of Mexico. Note the outrunner blocks (yellow
arrows). Compare with Figure 154, a seismic slice through the middle of this mass-transport deposit. Seismic data courtesy of
WesternGeco.
490 HENRY W. POSAMENTIER AND ROGER G. WALKER

Line 1

100 msec

Line 2

1
2
Line 3 3
1 km

FIG. 140.—Linear grooves at base of levees of slope channel shown in Figure 43. The events responsible for eroding these grooves
occurred just prior to levee construction, suggesting that these mass-transport events represent the earliest phases of a sea-level
lowstand. Seismic data courtesy of VeritasDGC.

Vertical section

Horizontal section

Debris “Clasts” Near Top of


Mass Transport Complex

52
0
m

FIG. 141.—Seismic reflection time slice as well as section view through a mass-transport complex with a single megaclast over 500
m wide highlighted. This clast is observed to be rafted near the top of the mass transport complex.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 491

Debris-flow
deposit

Base of
debris flow


FIG. 142.—Small-scale grooves and striations on top of sandstone
deposit at base of mass-transport complex, Borneo, Malaysia.

FIG. 143 (below).—Shelf-edge-detachment slump scars offshore


Indonesia. These slump scars likely represent the point of
detachment or staging area of sediments that traveled down
the slope, possibly transforming from slump to slide to flow
with increased distance from the shelf edge.
5 km Shelf edge
Shelf edge

Slump scar

1 km
492

Shelf edge Slump scar


A
B
Slope channel

Basin floor
C

Slump scar
HENRY W. POSAMENTIER AND ROGER G. WALKER

Scar Characteristics
Relief: 45 m
Width: ~ 16.4 km
Length: ~ 52 km
Area: 916 km2
Volume: 41.2 km3

FIG. 144.—Perspective A), dip-azimuth B), and section C) views across the slope of the eastern Gulf of Mexico illustrating massive slump/slide scars and evacuation
of massive amounts of material onto the basin floor to the south of this area. Seismic data courtesy of VeritasDGC.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 493

(i.e., corresponding to the magnitude of sea-level fall). Another


1 km factor that may play a role in destabilizing the slope at this time
is the shifting of oceanic currents that could accompany relative
falls of sea level. Such instability can commonly result in mass
failure in the mid- to upper-slope environment (Fig. 144) and
subsequent deposition of a mass-transport complex on the slope
and basin floor.
Ultimately, when river mouths are close enough to deliver
sediments directly to the outer shelf or upper slope, deep-water
turbidity currents become common across the slope and associ-
ated basin floor. During periods of relative sea-level fall, erosion
by incising rivers results in increased sediment delivery to the
shoreline, but more importantly increased sand-to-mud ratio of
sediments delivered there. Whether the sediments are delivered
directly from rivers by density underflow (i.e., hyperpycnal flow)
or are initially deposited and then later remobilized as slumps
transforming to slides and ultimately flows is not clear, and likely
both occur. At these times of relative sea-level fall, relatively
sand-rich flows tend to build relatively short leveed channels
transitioning into relatively large frontal splays.
When relative sea-level rise resumes, initially stationary shore-
lines with aggrading coastal deltas and plains (Fig. 57) and then
later transgressing shorelines result in a progressive decrease of
sands available for transport to the deep-water environment. At
these times coarser sands tend to be preferentially deposited
behind the shoreline within incised valleys, back-barrier lagoons,
and delta plains, producing gravity flows with potentially lower
sand-to-mud ratio late in the relative sea-level lowstand and
subsequent transgression. In response to this relative increase in
mud content within flow events, levee construction becomes
more efficient and leveed channels can extend significantly far-
ther basinward. The transition point is observed to shift signifi-
cantly farther seaward late in a sea-level cycle. This results in a
juxtaposition of frontal splays overlain by solitary leveed chan-
nels corresponding to early and late lowstand systems tract
times, respectively (Figs. 31, 74, 163, 164).
During subsequent periods of shoreline transgression, mini-
mal amounts of river-supplied sediments reach the shelf edge
and turbidite deposition largely ceases. Mass-transport deposi-
tion may continue at this time because of reequilibration of
residual oversteepened, upper-slope deposits associated with
lowstand shelf-edge delta deposition (Booth, 1979) or
oversteepening of canyon walls associated with lowstand canyon
or slope-channel cutting (Figs. 37, 165). A typical depositional
sequence within a canyon is illustrated in Figs. 37 and 165. In most
instances canyons are filled with relatively minor amounts of
channelized turbidites (Figs. 37–39) and an overwhelming amount
of mass-transport deposits that constitute the bulk of the canyon
FIG. 145.—Debris flow off flank of submarine mud volcano.
fill that forms during times when active turbidity flows have
ceased (Fig. 165).

The principal driver in deep-water sequence evolution is CONCLUSIONS


relative sea-level change—primarily the effect it has on character-
istics of the staging area, including sand-to-mud ratio, sediment The complexity of deep-water deposits is apparent from the
caliber, and depth of storm wave base. The stratigraphic, geomor- examples and discussions above. Large-scale, complex systems
phic, and sedimentologic expression of deep-water deposits is are particularly difficult to distill into simple models that can act
directly linked to the conditions at the shelf edge. It is there that as norms, predictors, and guides for future observations (Walker,
sediments are delivered by rivers and other processes from this volume).
hinterland areas. The locus of sedimentation, or depocenter, In the 1970s, little of this complexity was understood, and
shifts seaward and landward as a direct result of sea-level fall or models such as those of Normark (1970, 1978), Mutti and Ricci
rise, respectively. During sea-level fall, forced regression occurs Lucchi (1972), Mutti and Ghibaudo (1972), and Walker (1978)
on the shelf (Posamentier et al., 1992) and depocenters shift summarized observations from modern fans and ancient rocks,
rapidly toward the shelf edge. Even before the shoreline reaches and combined such observations into models based on modern
the shelf edge, instability of the upper slope has been exacerbated and ancient data (Walker, 1978). The models made little attempt
both by lowered wave base and by unloading of a wedge of water to incorporate data from turbidity-current processes (laboratory
494

MTC
MTC lobes
channel

Line 1

Levees
Line 1
Deflated MTC lobe
HENRY W. POSAMENTIER AND ROGER G. WALKER

MTC
5 km

FIG. 146.—Seascape of the lower slope and basin floor, DeSoto Canyon area, eastern Gulf of Mexico. Note the varying types of mass-transport deposits, including a
leveed mass-transport deposit, a lobate mass-transport deposit characterized by positive relief, and a flattened or deflated mass-transport deposit showing no
positive relief and characterized only by a debris field. Seismic data courtesy of VeritasDGC.
A Knickpoint • Channel depth = ~ 26 m
• Channel width = ~ 1.2 km

Debr
is-flo
w ch
anne
l

Turbidity-current channel

B C
DEEP-WATER TURBIDITES AND SUBMARINE FANS

five
five km
km

FIG. 147.—A) Dip-magnitude, B) curvature, and C) time-structure maps of a debris-flow channel. This channel is characterized by low sinuosity and the presence of
a well-defined knickpoint at its head. Seismic data courtesy of WesternGeco.
495
496 HENRY W. POSAMENTIER AND ROGER G. WALKER

A B Line 1
five km

MTC Lobes

n
tio
rec
Di
l w
Fo
Structural Ridge

South North FIG. 148.—Illuminated sea floor based on 3D seismic data


A) and interpretation of mass-transport deposits B) on
C five km
the basin floor of the Makassar Strait, Indonesia. Sec-
tion view C) through several shallowly buried mass-
transport lobes of various age. Note that these elon-
gate lobes represent flow complexes imbedded within
a low-amplitude seismic reflection package sugges-
tive of hemipelagic to pelagic sedimentation. Because
of the draping effect of the hemipelagic and pelagic
sediments, these buried lobes all have sea-floor ex-
pression. Seismic data courtesy of WesternGeco.

FIG. 149.—Seismic reflection horizon slice showing an


“opportunistic” mass-transport deposit within an
earlier-formed turbidity-current leveed channel.
Seismic data courtesy of WesternGeco.

one km
B

Flo
wd
irec
tion

one km
South North
one km
A

Lobe 4
Slice level

100
msec
Lobe 1 Lobe 3

B
five km
DEEP-WATER TURBIDITES AND SUBMARINE FANS

Lobe 3
Flow direction

Lobe 2

Lobe 4

FIG. 150.—Seismic reflection cross-section A) and plan-view B) images illustrating a succession of amalgamating mass-transport deposits. Note the typical chaotic to
transparent reflection character within these deposits, suggesting a severely disrupted and discontinuous stratigraphic architecture. Seismic data courtesy of
WesternGeco.
497
Compressional ridges 498
A

five km

one km
100 msec
B
Flow direction

Flow
Flow Direction
Direction
Terminal wall

C D
HENRY W. POSAMENTIER AND ROGER G. WALKER

15°

100 msec

one km
FIG. 151.—Seismic reflection plan view A) and section views B–D) through a mass-transport complex, Gulf of Mexico. The transverse lineaments are the plan-view
expression of thrust faults caused by the mass flow abutting against the terminal wall. Thrust faults are measured at approximately a 15 degree dip. Seismic data
courtesy of WesternGeco.
A

5 km C
(approximate)

Event 3
DEEP-WATER TURBIDITES AND SUBMARINE FANS

Event 1

Mass-transport complex

Toe-thrust faults
Event 2
Lateral
compression
FIG. 152.—A, B) Oblique and C) section views through an amalgamated mass-transport complex. At least three event units can be recognized. Internal deformation in
the form of compression-associated low-angle thrust faults can be observed within events 2 and 3 (B). Both toe thrusts as well as thrusts associated with lateral
compression caused by later flows are seen. Seismic data courtesy of WesternGeco.
499
500

2 km

Nigeria

2 km

Indonesia
HENRY W. POSAMENTIER AND ROGER G. WALKER

2 km

Gulf of Mexico
FIG. 153.—Cross-section view through three mass-transport deposits. Each is characterized by chaotic and contorted seismic facies pattern. Top line, seismic data
courtesy of VeritasDGC. Middle and lower lines, seismic data courtesy of WesternGeco.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 501

one km
FIG. 154.—Seismic horizon slice through mass-transport deposit characterized by convolute bedding, basin floor of the eastern Gulf of
Mexico. Compare with Figure 139, a seismic slice at the base of this mass-transport deposit. Seismic data courtesy of WesternGeco.

or modern ocean observations), and thus remained observational the seismic data revealed so well the geometrical relationships
rather than process based. In retrospect, we note that the Walker between elements, it became possible to infer and better under-
(1978) model showed a very simplistic and commonly incorrect stand the processes that gave rise to the various depositional
distribution of sand and mud on the fan, and also ignored the geometries—processes such as channel meandering and avul-
effects of different grain-size distributions, tectonic settings, and sion, and lobe development by crevassing, overbank spill, or flow
relative sea-level fluctuations. expansion at the ends of channels.
Relative sea-level fluctuations were built into the models of Thus any attempts at modeling of turbidite systems must
Mutti (1985), Vail et al. (1977), and Posamentier and Vail (1985). incorporate information of four different types and scales: pro-
This model essentially combined three aspects of older Mutti cesses (on the flume experimental scale as well as the scale of
models into one evolving model, with sandy detached lobes modern fans), 3-D seismic studies, observations of ancient rocks,
forming during lowstands of relative sea level. During relative and numerical models. There is still a rather incomplete linkage
sea-level rise, the fan evolved into one where the channels were between these different sources of data, and we have tried to
attached to smaller depositional lobes. At highstand, with a much show how some of these linkages might be made. Our conclu-
reduced sediment supply, either condensed sections or finer- sions are based in fact as much as possible but are nevertheless
grained channels and levee formed, with little or no lobe forma- somewhat process and conceptual in flavor. Future fan models
tion. will evolve as more work is done relating experimental studies to
The 1980s marked the time when geophysical observations ancient rocks and relating the geometry of large outcrops to 3-D
became important, particularly side-scan sonar and shallow- seismic data derived both from modern fans and Cretaceous–
penetration seismic profiles across fan surfaces. This information Tertiary turbidite systems. As the seismic studies evolve, so the
was very different in nature and scale from the rock observations large-scale processes of fan development will be better under-
and poorer-quality side-scan data from the 1970s. Because the stood, and the future conceptual basis for modeling will be better
data were so different—for example, data on the meandering established.
channels from the Amazon Fan (Damuth et al., 1982a; Damuth et In this paper, we have integrated facies with depositional
al., 1982b) and the data on channel, levee, and large-scale slump elements and then discussed modern and subsurface 3-D seis-
from the Rhône Fan (Droz and Bellaiche, 1985), it was difficult to mic examples to show how the depositional facies fit into the
combine older models with this new data—in fact, there are no various depositional systems. Because in most instances ground
published attempts to do so. truth is lacking in studies based on 3D seismic, we have inte-
At around the same time, 3-D seismic data from Tertiary and grated information from outcrop examples to help provide
Modern fans was becoming available to oil companies, but few of insights to assist in interpreting the 3-D seismic images (for
these studies made their way into the public literature until the example, the San Clemente and Ross outcrop examples help to
1990s. The seismic data revealed superb three-dimensional rela- calibrate the seismic images of meander loops; Figs. 51, 71, 72).
tionships between all of the depositional elements, but again it Rather than create a model that would serve as a template, we
was difficult to relate seismic-scale data to outcrop data. With the have focused on first principles based largely on process sedi-
advent of seismic geomorphology (Posamentier and Kolla, 2003a) mentology. In this way, we leave it to geoscientists to build their
the integration of such observations was made possible. Because own models using depositional elements as building blocks in
A B
B C
C 502
A A B A
A A'
A'

Terminal wall

Slice
C'
C' level
B'
B'

1 km

A' one km C
A'

Leveed channel
Turbidite
frontal splay

Mass-transport complex

D B B'
B B'

100 m
HENRY W. POSAMENTIER AND ROGER G. WALKER

E C
C C'
C'

one km

FIG. 155.—Example of the terminus of a mass-transport deposit, eastern Gulf of Mexico. This deposit shows evidence of having been transported across a decollement
surface that likely was located within a condensed section at the base of a frontal-splay complex. In response to compression against a terminal wall, internal
deformation in the form of thrust faulting occurred. The plan-view A) as well as the section views B–E) the clearly show the mass-transport unit entraining earlier-
deposited, sand-prone leveed-channel and frontal-splay deposits. Seismic data courtesy of WesternGeco.
A B Transport
direction

Small thrust faults

C
D
DEEP-WATER TURBIDITES AND SUBMARINE FANS

Transport
direction

Transport
direction

FIG. 156.—Seismic time slice in the coherence A) and amplitude B) domains illustrating the plan view and oblique view C) expression of thrust faults within mass-transport
deposits. D) Seismic reflection profile across mass transport deposits characterized by low-angle thrust faults. Seismic data courtesy of WesternGeco.
503
504 HENRY W. POSAMENTIER AND ROGER G. WALKER

can be characterized by the full range of sand habitats, insofar as


they comprise, in microcosm, slope and basin-floor physiogra-
phy. Figures 166–168 illustrate and summarize the various
significant sand habitats on the basin floor. Figures 169 and 170
illustrate the significant sand habitats on the slope.
The role of grain-size distribution within a flow as well as in
a succession of flows is critical to understanding the evolution of
depositional elements, especially those on the basin floor. The
more mud in the system, the farther seaward the leveed-channel
depositional element extends. As flows travel seaward they are
preferentially impoverished of their mud content through con-
tinual spillover of the upper parts of the flows into the overbank
environment. At the point where the levees are low enough to
allow sand from the higher-density, lower part of the flow to spill
over the levees, the system morphology transitions from leveed
channel to frontal splay. The frontal splays are commonly chan-
nelized in their proximal parts, but the channels are shallow and
have small levees. Channel paths probably switch rapidly. In the
distal parts of the frontal splays, flows become even less channel-
ized and more sheet-like.
Mass-transport deposits, largely debrites, are common com-
ponents of deep-water environments (Fig. 133). These deposits,
Trench
which are largely mud prone, commonly originate in mid-slope
to upper-slope environments. They commonly directly overlie
erosional surfaces characterized by long grooves and striations.
Near their termini as well as along their margins, they are
Thrust faults characterized by internal compressive deformation in the form of
imbricate thrust faults. Mass-transport complexes likely are char-
acterized by complex rheology that reflects frequent transitions
from turbulent to laminar to “plug” flow within a single flow
event.
In summary, we urge caution in adopting overly simple facies
models in deep-water environments. The deep water is a poten-
tially complex depositional setting. The degree of complexity of
the facies model desired should be dependent on the quality of
data available and should be built upon depositional elements
observed and inferred. However, as the geoscientist well knows,
FIG. 157.—Small-scale analog for mass-transport deposits shown the devil is in the details!
in Figures 134, 135, and 155. The snow shovel slides on a hard
base, pushing snow before it. The semi-rigid snow pack ACKNOWLEDGMENTS
deforms internally predominantly by low-angle thrust fault-
ing as the mass slides on the underlying decollement surface The authors acknowledge Anadarko Petroleum Corporation
and is compressed against the snow pack before it. The trench, for permission to publish. Permission to publish seismic data
which is formed in the wake of the shovel, is characterized by from VeritasDGC, Western Geophysical, PGS Geophysical is
steep walls produced by shearing of the flowing mass. gratefully acknowledged. We also wish to thank the various
colleagues with whom we have had lengthy conversations about
the world of deep-water sedimentation through the years. These
space and time. On the basis of a sound understanding of local include V. Kolla, P. Weimer, W. Normark, H. DeV. Wickens, B.
physiography, sediment flux, and sediment caliber, integrated Kneller, and W. Morris. This paper benefited from and was
with process sedimentology, predictive facies models can be significantly improved thanks to the comprehensive reviews of
constructed. Lateral as well as vertical facies successions can be Todd Greene and Octavian Catuneaunu.
predicted.
Sediment accumulates in staging areas, moves down can- REFERENCES
yons or slope channels, and ultimately travels onto the basin
floor through leveed channels and frontal splays. Principal ABREU, V., SULLIVAN, M., PIRMEZ, C., AND MOHRIG, D., 2003, Lateral accretion
deep-water sand habitats include slope-channel complexes, packages (LAPs): an important reservoir element in deep water
basin-floor channel fills, crevasse splays and sediment waves in sinuous channels: Marine and Petroleum Geology, v. 20, p. 631–648.
overbank settings, and frontal splays. Canyons also can be sites BALLANCE, P.F., 1964, The sedimentology of the Waitemata Group in the
of confined leveed channel deposits, though the bulk of canyon Takapuna section, Auckland: New Zealand Journal of Geology and
fill commonly is mud prone. To a lesser extent, sand deposition Geophysics, v. 7, p. 466–499.
also occurs within levees in response to repeated spillover BEAUBOEUF, R.T., AND FRIEDMANN, S.J., 2000, High-resolution seismic/
across levee crests. Where slopes are characterized by high sequence stratigraphic framework for the evolution of Pleistocene
rugosity, such as areas underlain by mobile salt or complex toe- intra slope basins, Western Gulf of Mexico: depositional models and
thrust ridges, deposition of a broad range of depositional ele- reservoir analogs, in Weimer, P., Slatt, R.M., Coleman, J., Rosen, N.C.,
ments can occur within relatively small areas. Intraslope basins Nelson, H., Bouma, A.H., Styzen, M.J., and Lawrence, D.T., eds.,
A B

C D

Slice level
DEEP-WATER TURBIDITES AND SUBMARINE FANS

FIG. 158.—Seismic time slice and section views A, B, D) of mass-


transport deposits characterized by thrust faults induced by
compressional stress regime. The interpreted thrust faults C)
formed as a result of “country rock” being “pushed” and com-
pressed by stress directed from the right by a mass flow E).
505
Distance traveled = 6.6 km 506
[Assumption: 30 faults @ 220 m/fault]

Terminal wall

Restoration of fault results in 220 m of lengthening Distance Traveled = 0 km

100 msec
HENRY W. POSAMENTIER AND ROGER G. WALKER

Terminal wall
one km

FIG. 159.—Seismic profile through mass-transport deposit shown in Figure 149. Approximate palinspastic reconstruction results in 220 m of lengthening for each thrust
fault. With approximately 30 faults along an axial profile within the study area, restoration results in extension of approximately 6.6 km at the up-system limit of
the deposit, whereas it can be assumed that the sediments near the terminal wall would have moved only minimally. Seismic data courtesy of WesternGeco.
A Turbidite channel B

Lateral
compression

MTD Transport
margin direction

C
DEEP-WATER TURBIDITES AND SUBMARINE FANS

Lateral
compression

FIG. 160.—A) Illuminated horizon at the top of a mass-transport deposit illustrating a lateral compressional bulge. B, C) Seismic reflection profile illustrating lateral
compression associated with a later phase of mass transport.
507
508 HENRY W. POSAMENTIER AND ROGER G. WALKER

A Transition point

Transition point

Onlap of frontal splay

Transport
direction

FIG. 161.—Influence on turbidites by rugosity atop mass-transport deposits. A) Seismic time slice that shows the transition point
of a frontal splay, with extensive frontal-splay deposits seaward of that location. B) Seismic section that illustrates the onlap
of frontal-splay deposits against a bathymetric high associated with the irregular top of a mass-transport complex. Seismic data
courtesy of WesternGeco.

Deep-Water Reservoirs of the World: Gulf Coast Section SEPM, 20th BRAMI, T.R., PIRMEZ, C., ARCHIE, C., HEERALAL, S., AND HOLMAN, K.L., 2000,
Annual Research Conference, p. 40–60. Late Pleistocene deep-water stratigraphy and depositional processes,
BOOTH, J.S., 1979, Recent history of mass-wasting on the upper continental offshore Trinidad and Tobago, in Weimer, P., Slatt, R.M, Coleman, J.,
slopes, northern Gulf of Mexico, as interpreted from the consolidation Rosen, N.C., Nelson, H., Bouma, A.H., Styzen, M.J., and Lawrence,
states of the sediment, in Doyle, L.J., and Pilkey, O.H., eds., Geology D.T., eds., Deep-Water Reservoirs of the World: Gulf Coast Section
of Continental Slopes: SEPM, Special Publication 27, p. 153–165. SEPM Foundation, 20th Annual Research Conference, p. 104–115.
BOUMA, A.H., 1962, Sedimentology of Some Flysch Deposits; A Graphic BRUHN, C.H.L., AND WALKER, R.G., 1995, High-resolution stratigraphy and
Approach to Facies Interpretation: Amsterdam, Elsevier, 168 p. sedimentary evolution of coarse-grained canyon-filling turbidites
BOUMA, A.H., NORMARK, W.R., AND BARNES, N.E., eds., 1985, Submarine Fans from the Upper Cretaceous transgressive megasequence, Campos
and Related Turbidite Systems: New York, Springer-Verlag, 351 p. Basin, Brazil: Journal of Sedimentary Research, v. B65, p. 426–442.
A B

C Scarp relief up to 22 m
D
Relief up to 32 m
DEEP-WATER TURBIDITES AND SUBMARINE FANS

one km

FIG. 162.—Rugose upper surface of a mass-transport deposit. Small-scale accommodation atop the mass-transport deposits shows no apparent influence on
subsequent turbidite deposition.
509
510

B
HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 163.—A) Seismic reflection profile across deep-water deposits on the basin floor of the Makassar Strait, Indonesia, illustrating a stratigraphic succession with mass-
transport deposits at the base, overlain by frontal-splay deposits, leveed-channel deposits, a further mass-transport deposit and ultimately a condensed-section
deposit. B) Idealized cross section and well logs through a deep-water depositional sequence (after Posamentier and Kolla, 2003a). Seismic data courtesy of
WesternGeco.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 511

Relative sea level


High

Lowstand interval

Low Time

Condensed section
Interval that shoreline is
located near the shelf edge
(turbidity currents dominate)
Interval of upper-slope
instability (mass-transport dominates)
Leveed channels dominate
(relatively low sand:mud)
Frontal splays dominate
(relatively high sand:mud)

Interval of upper-slope instability Condensed


(mass-transport dominates) section

FIG. 164.—Schematic depiction of sediment transport events associated with relative sea-level change. As sea level begins its fall,
lowered wave base results in slope disequilibrium conditions, which favor mass-transport events. Once sea level has lowered
to the point where river mouths are in close proximity to the shelf edge, direct and indirect delivery of turbidites to the slope
and basin floor is facilitated. During the early stages of this process, associated with the interval of relative sea-level fall, shelf
valleys are incised and canyons can form. This results in sediment bypass of the shelf, which favors delivery of a relatively sand-
prone sediment load to the deep water. During the late stages of sea-level lowstand, when sea level is slowly rising, sediments
(preferentially coarse-grained sediment) tend to be trapped within earlier-formed incised valleys, resulting in muddier
turbidites at that time. This progression favors an evolution from sand-rich frontal splays to mud-rich isolated leveed channels.
Finally rapid sea-level rise again is associated with slope disequilibrium and deposition of mass-transport deposits. When sea
level finally stabilizes, background deposition of hemipelagic and pelagic sediments dominates in the deep-water environ-
ment, forming a condensed section.

BRUHN, C.H.L., AND WALKER, R.G., 1997, Internal architecture and sedi- M.J., and Lawrence, D.T., eds., Deep-Water Reservoirs of the World:
mentary evolution of coarse-grained turbidite channel–levee com- Gulf Coast Section SEPM Foundation, 20th Annual Research Confer-
plexes, Early Eocene Regencia Canyon, Espirito Santo Basin, Brazil: ence, p. 127–150.
Sedimentology, v. 44, p. 17–46. CLIFTON, H.E., 1981, Submarine canyon deposits, Point Lobos, California,
CAMACHO, H., BUSBY, C.J., AND KNELLER, B., 2002, A new depositional model in Frizzel, E., ed., Upper Cretaceous and Paleocene Turbidites, Cen-
for the classical turbidite locality at San Clemente State Beach, Califor- tral California Coast: Pacific Section SEPM, Guidebook to Field Trip
nia: American Association of Petroleum Geologists, Bulletin, v. 86, p. no. 6, p. 79–92.
1543–1560. CLIFTON, H.E., 1984, Sedimentation units in stratified resedimented con-
CAMPION, K.M., SPRAGUE, A.R., MOHRIG, D., LOVELL, R.W., DRZEWIECKI, P.A., glomerate, Paleocene submarine canyon fill, Point Lobos, California,
SULLIVAN, M.D., ARDILL, J.A., JENSEN, G.N. AND SICKAFOOSE, D.K., 2000, in Koster, E.H., and Steel, R.J., eds., Sedimentology of Gravels and
Outcrop expression of confined channel complexes, in Weimer, P., Conglomerates: Canadian Society of Petroleum Geologists, Memoir
Slatt, R.M, Coleman, J., Rosen, N.C., Nelson, H., Bouma, A.H., Styzen, 10, p. 429–-441.
512
10 km

450 m

1170 m
Mass-transport deposits

18°

Seismic section
HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 165.—Seismic section across Mississippi canyon, along with perspective shaded relief. Seismic data proprietary to PGS Marine Geophysical NSA.
Mud plug

Blocky to fining upward

50 m
one km Amalgamated
channel fill
DEEP-WATER TURBIDITES AND SUBMARINE FANS

FIG. 166.—Basin-floor leveed-channel sand habitats—channel. Seismic data courtesy of WesternGeco.


513
514

50 m

one km

Crevasse splay
Sediment waves
Thin-bedded overbank
HENRY W. POSAMENTIER AND ROGER G. WALKER

FIG. 167.—Basin-floor leveed-channel sand habitats—levee. Seismic data courtesy of WesternGeco.


50 m
one km

Frontal splay
Frontal splay
DEEP-WATER TURBIDITES AND SUBMARINE FANS

FIG. 168.—Basin-floor frontal-splay sand habitats. Seismic data courtesy of WesternGeco.


515
516

Onlap at channel wall with shale drape


HENRY W. POSAMENTIER AND ROGER G. WALKER

Slope-channel lateral-accretion package

Slope-channel amalgamated fill


FIG. 169.—Slope-channel sand habitats. Seismic data courtesy of VeritasDGC.
Canyon filled with predominantly mass transport deposits

one km

one km

Partially preserved channel


(erosional remnant) Coarse-grained canyon-axis fill
DEEP-WATER TURBIDITES AND SUBMARINE FANS

one km

Axial leveed channel Canyon-fill onlap onto canyon wall

FIG. 170.—Slope-canyon sand habitats. Seismic data proprietary to PGS Marine Geophysical NSA.
517
518 HENRY W. POSAMENTIER AND ROGER G. WALKER

CROWELL, J.C., 1957, Origin of pebbly mudstones: Geological Society of Reservoirs of the World: Gulf Coast Section SEPM Foundation, 20th
America, Bulletin, v. 68, p. 993–1009. Annual Research Conference, p. 402–421.
DAMUTH, J.E., KOWSMANN, R.O., FLOOD, R.D., BELDERSON, R.H., AND GORINI, KENYON, N.H., AND MILLINGTON, J., 1995, Contrasting deep-sea deposi-
M.A., 1983a, Age relationships of distributary channels on Amazon tional systems in the Bering Sea, in Pickering, K.T., Hiscott, R.N.,
deep-sea fan: implications for fan growth pattern: Geology, v. 11, p. Kenyon, N.H., Ricci Lucchi, F., and Smith, R.D.A., eds., Atlas of Deep
470–473. Water environments: London, Chapman & Hall, p. 196–202.
DAMUTH, J.E., KOLLA, V., FLOOD, R.D., KOWSMANN, R.O., GORINI, M.A., AND KNELLER, B.C., 1995, Beyond the turbidite paradigm: physical models
BELDERSON, R.H., 1983b, Distributary channel meandering and bifur- for deposition of turbidites and their implications for reservoir
cation patterns on the Amazon deep-sea fan as revealed by long- prediction, in Hartley, A.J., and Prosser, D.J., eds., Characterisation
range side-scan sonar (GLORIA): Geology, v. 11, p. 94–98. of Deep Marine Clastic Systems: Geological Society of London,
DROZ, L., AND BELLAICHE, G., 1985, Rhône deep-sea fan: morphostructure Special Publication 94, p. 31–49.
and growth pattern: American Association of Petroleum Geologists, KOLLA, V., AND COUMES, F., 1987, Morphology, internal structure, seis-
Bulletin, v. 69, p. 460–479. mic stratigraphy and sedimentation of Indus Fan: American Asso-
DZULYNSKI, S., KSIAZKIEWICZ, M., AND KUENEN, P.H., 1959, Turbidites in ciation of Petroleum Geologists, Bulletin, v. 71, p. 650–677.
flysch of the Polish Carpathian Mountains: Geological Society of KOLLA, V., BOURGES, P., URRUTY, J.M., AND SAFA, P., 2001, Evolution of
America, Bulletin, v. 70, p. 1089–1118. deepwater Tertiary sinuous channels offshore, Angola (West Af-
ELMORE, R.D., PILKEY, O.H., CLEARY, W.J., AND CURRAN, H.A., 1979, Black rica) and implications to reservoir architecture: American Associa-
Shell turbidite, Hatteras Abyssal Plain, western Atlantic Ocean: tion of Petroleum Geologists, Bulletin v. 85, p. 1373–1405.
Geological Society of America, Bulletin, v. 90, p. 1165–1176. KOMAR, P.D., 1971, Hydraulic jumps in turbidity currents: Geological
GHIBAUDO, G., AND VANZ, V., 1987, Proposta di classificazione delle facies Society America, Bulletin, v. 82, p. 1477–1488.
“torbiditiche” e di un metodo practico per la loro descrizione sul KUENEN, P.H., AND MIGLIORINI, C.I., 1950, Turbidity currents as a cause
terreno: Giornale di Geologia, v. 49, p. 31/43. of graded bedding: Journal of Geology, v. 58, p. 91–127.
GOODWIN, R.H., AND PRIOR, D.B., 1989, Geometry and depositional se- LIEN, T., WALKER, R.G., AND MARTINSEN, O.J., 2003, Turbidites in the
quences of the Mississippi Canyon, Gulf of Mexico: Journal of Upper Carboniferous Ross Formation, western Ireland: recon-
Sedimentary Petrology, v. 59, p. 318–329. struction of a channel and spillover system: Sedimentology, v. 50,
GREGORY, M.R., 1966, Sedimentary features and penecontemporaneous p. 113-148.
slumping in the Waitemata Group, Whangaparaoa Peninsula, north LINK, M.H., SQUIRES, R.L., AND COLBURN, I.P., 1984, Slope and deep-sea
Auckland, New Zealand: New Zealand Journal of Geology and fan facies and paleogeography of Upper Cretaceous Chatsworth
Geophysics, v. 12, p. 248–282. Formation, Simi Hills, California: American Association of Petro-
GRIGGS, G.B., AND KULM, L.D., 1970, Sedimentation in Cascadia deep-sea leum Geologists, Bulletin, v. 68, p. 850–873.
channel: Geological Society of America, Bulletin, v. 81, p. 1361–1384. LOWE, D.R., 1975, Water escape structures in coarse-grained sedi-
HACKBARTH, C.J., AND SHAW, R.D., 1994, Morphology and stratigraphy of ments: Sedimentology, v. 22, p. 157–204.
a mid-Pleistocene turbidite leveed channel from seismic, core and LOWE, D.R., 1979, Stratigraphy and sedimentology of the Pigeon point
log data, northeastern Gulf of Mexico, in Weimer, P., Bouma, A.H., Formation, San Mateo County, California, in Nilsen, T.H., and
and Perkins, B.F., eds., Submarine Fans and Turbidite Systems: Gulf Brabb, E.E., eds., Geology of the Santa Cruz Mountains, California:
Coast Section SEPM Foundation, 15th Annual Research Conference, Geological Society of America, Cordilleran Section, Field Trip
p. 127–133. Guidebook, p. 17–29.
H ARMS, J.C., AND F AHNESTOCK, K., 1965, Stratification, flow phenom- LOWE, D.R., 1982, Sediment gravity flows: II. Depositional models with
ena and bed forms (with an example from the Rio Grande), in special reference to the deposits of high-density turbidity currents:
Middleton, G.V., ed., Primary Sedimentary Structures and Their Journal of Sedimentary Petrology, v. 52, p. 279–297.
Hydrodynamic Interpretation: SEPM , Special publication 12, p. MANLEY, L., AND FLOOD, R.D., 1998, Cyclic sediment deposition within
84–115. Amazon deep-sea fan: American Association of Petroleum Geolo-
HEEZEN, B.C., 1956, Corrientes de turbidez del Rio Magdalena: Sociedad gists, Bulletin, v. 72, p. 912–925.
Geográfica de Colombia, Boletín, v. 51/52, p. 135–143. MARTINSEN, O.J., LIEN, T., AND WALKER, R.G., 2000, Upper Carboniferous
HEEZEN, B.C., MENZIES, R.J., SCHNEIDER, E.D., EWING W.M., AND GRANELLI, deep-water sediments, western Ireland: analogs to passive margin
N.C.L., 1964, Congo submarine canyon: American Association of plays, in Weimer, P., Slatt, R.M., Coleman, J., Rosen, N.C., Nelson,
Petroleum Geologists, Bulletin, v. 48, p. 1126–1149. H., Bouma, A.H., Styzen, M.J., and Lawrence, D.T., eds., Deep
HISCOTT, R.N., HALL, F.R., AND PIRMEZ, C., 1997, Turbidity-current overspill Water Reservoirs of the World: SEPM, Gulf Coast Section, 20th
from the Amazon Channel: texture of the silt/sand load, paleoflow Annual Bob F. Perkins Research Conference, Houston, p. 533–555
from anisotropy of magnetic susceptibility and implications for (available only on CD).
flow processes, in Flood, R.D., Piper, D.J.W., Klaus, A., and Peterson, MASLIN, M., MIKKELSEN, N., VILELA, C., AND HAQ, B.U., 1998, Sea-level
L.C., eds., Proceedings of the Ocean Drilling Program: Scientific and gas-hydrate–controlled catastrophic sediment failures of the
Results, v. 155, p. 53–78. Amazon fan: Geology, v. 26, p. 1107–1110.
HURST, A., CARTWRIGHT, J.A., DURANTI, D., HUUSE, M., AND NELSON, M., MAYALL, M., AND STEWART, I., 2000, The architecture of turbidite slope
2005, Sand injectites: an emerging global play in deep-water clastic channels, in Weimer, P., Slatt, R.M, Coleman, J., Rosen, N.C.,
environments, in Doré, A.G., and Vining, B.A., eds., Petroleum Nelson, H., Bouma, A.H., Styzen, M.J., and Lawrence, D.T., eds.,
Geology: Northwest Europe and Global Perspectives: Geological Deep-Water Reservoirs of the World: Gulf Coast Section SEPM
Society of London, Proceedings of the 6th Petroleum Geology Con- Foundation, 20th Annual Research Conference, p. 578–586.
ference, p. 133–144. MIGEON, S., SAVOYE, B., AND FAUGERES, J.-C., 2000, Quaternary develop-
JENNETTE, D.C., GARFIELD, T.R., MOHRIG, D.C., AND CAYLEY, G.T., 2000, The ment of migrating sediment waves in the Var deep-sea fan: distri-
interaction of shelf accommodation, sediment supply and sealevel bution, growth pattern and implication for levee evolution: Sedi-
in controlling the facies, architecture and sequence stacking pat- mentary Geology, v. 133, p. 265–293.
terns of the Tay and Forties/Sele basin-floor fans, Central North Sea, MOHRIG, D., WHIPPLE, K.X., HONDZO, M., ELLIS, C., AND PARKER, G., 1998,
in Weimer, P., Slatt, R.M, Coleman, J., Rosen, N.C., Nelson, H., Hydroplaning of subaqueous debris flows: Geological Society of
Bouma, A.H., Styzen, M.J., and Lawrence, D.T., eds., Deep-Water America, Bulletin, v. 110, p. 387–394.
DEEP-WATER TURBIDITES AND SUBMARINE FANS 519

MULDER, T., AND SYVITSKI, J.P.M., 1995, Turbidity currents generated at PIRMEZ, C., HISCOTT, R.N., AND KRONEN, J.K., 1997, Sandy turbidite succes-
river mouths during exceptional discharges to the world oceans: sions at the base of channel–levee systems of the Amazon Fan re-
Journal of Geology, v. 103, p. 285–299. vealed by FMS logs and cores: unraveling the facies architecture of
MULDER, T., SYVITSKI, J.P.M., AND SKENE, K.I., 1998, Modeling of erosion large submarine fans, in Flood, R.D., Piper, D.J.W., Klaus, A., and
and deposition by turbidity currents generated at river mouths: Peterson, L.C., eds., Proceedings of the Ocean Drilling Program,
Journal of Sedimentary Research, v. 68, p. 124–137. Scientific Results, v. 155, p. 7–22.
MUTTI, E., 1985, Turbidite systems and their relations to depositional PLINT, A.G., AND NUMMEDAL, D., 1998, The falling stage systems tract:
sequences, in ZUFFA, G.G., ed., Provenance of Arenites: NATO-ASI recognition and importance in sequence stratigraphic analysis, in
Series, Dordrecht, The Netherlands, Reidel, p. 65–93. Gawthorpe, R.L.G., and Hunt, E., eds., Sedimentary Responses to
MUTTI, E., AND GHIBAUDO, G., 1972, Un esempio di torbiditi di conoide Forced Regression: Geological Society of London, Special Publica-
sottomarina esterna: le Arenario di San Salvatore (formazione di tion.
Bobbio, Miocene) nell’Appennino di Piacenza: Memorie POSAMENTIER, H.W., 2003a, A linked shelf edge delta and slope channel
dell’Accademia delle Scienze di Torino, Classe di Scienze Fisiche, turbidite system: 3D seismic case study from the eastern Gulf of
Matematiche e Naturale, Serie 4a, no. 16, p. 1–41. Mexico, in Roberts, H.H., Rosen, N.C., Fillon, R.H., and Anderson,
MUTTI, E., AND NORMARK, W.R., 1991, An integrated approach to the study J.B., eds., Shelf Margin Deltas and Linked Down Slope Petroleum
of turbidite systems, in Weimer, P., and Link, M.H., eds., Seismic Systems: Global Significance and Future Exploration Potential: Gulf
Facies and Sedimentary Processes of Submarine Fans and Turbidite Coast Section of SEPM Foundation, Proceedings of the 23rd Annual
Systems: New York, Springer-Verlag, p. 75–106. Bob F. Perkins Research Conference, p. 115–134.
MUTTI, E., AND RICCI LUCCHI, F., 1972, Le torbiditi dell’Appennino POSAMENTIER, H.W., 2003b, Depositional elements associated with a basin
settentrionale: introduzioni all’analisi de facies: Società Geologica floor channel–levee system: case study from the Gulf of Mexico:
Italiana Memorie, v. 11, p. 161–199. Marine and Petroleum Geology, v. 20, p. 677–690.
MUTTI, E., STEFFENS, G.S., PIRMEZ, C., ORLANDO, M., AND ROBERTS, D., EDS., POSAMENTIER, H.W., AND ALLEN, G.P., 1999, Siliciclastic Sequence Stratigra-
2003, Thematic set: Turbidites: models and problems: Marine and phy—Concepts and Applications: SEPM, Concepts in Sedimentology
Petroleum Geology, v. 20, p. 523–933. and Paleontology, no. 7, 210 p.
NELSON, C.H., GOLDFINGER, C., JOHNSON, J.E., AND DUNHILL, G., 2000, Varia- POSAMENTIER, H.W., ALLEN, G.P., JAMES, D.P., AND TESSON, M., 1992, Forced
tion of modern turbidite systems along the subduction zone margin regressions in a sequence stratigraphic framework: concepts, ex-
of Cascadia Basin and implications for turbidite reservoir beds, in amples, and exploration significance: American Association of Petro-
Weimer, P., Slatt, R.M., Coleman, J., Rosen, N.C., Nelson, H., Bouma, leum Geologists, Bulletin, v. 76, p. 1687–1709.
A.H., Styzen, M.J. and Lawrence, D.T., eds., Deep Water Reservoirs of POSAMENTIER, H.W., ERSKINE, R.D., AND MITCHUM, R.M., JR., 1991, Models for
the World: SEPM, Gulf Coast Section, 20th Annual Bob F. Perkins submarine fan deposition within a sequence stratigraphic frame-
Research Conference, Houston, p. 714–738 (available only on CD). work, in Weimer, P., and Link, M.H., eds., Seismic Facies and Sedi-
NILSEN, T.H., 1980, Modern and ancient submarine fans: discussion of mentary Processes of Submarine Fans and Turbidite Systems: New
papers by R.G. Walker and W.R. Normark: American Association of York, Springer-Verlag, p. 127–136.
Petroleum Geologists, Bulletin, v. 64, p. 1094–1101. POSAMENTIER, H.W., JERVEY, M.T., AND VAIL, P.R., 1988, Eustatic controls on
NORMARK, W.R., 1970, Growth patterns of deep-sea fans: American Asso- clastic deposition I—conceptual framework, in Wilgus, C.K., Hastings,
ciation of Petroleum Geologists, Bulletin, v. 54, p. 2170–2195. B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., and Van
PEAKALL, J., MCCAFFREY, W.D., KNELLER, B.C., STELTING, C.E., MCHARGUE, Wagoner, J.C., eds., Sea Level Changes: An Integrated Approach:
T.R., AND SCHWELLER, W.J., 2000, A process model for the evolution of SEPM, Special Publication 42, p. 109–124.
submarine fan channels: Implications for sedimentary architecture, in POSAMENTIER, H.W., AND KOLLA, V., 2003a, Seismic geomorphology and
Bouma, A.H., and Stone, C.G., eds., Fine-Grained Turbidite Systems: stratigraphy of depositional elements in deep-water settings: Journal
American Association of Petroleum Geologists, Memoir 72 and SEPM, of Sedimentary Research, v. 73, p. 367–388.
Special Publication 68, p. 73–88. POSAMENTIER, H.W., AND KOLLA, V., 2003b, Anatomy of a deep-water
PICKERING, K.T., STOW, D., WATSON, M., AND HISCOTT, R.N., 1986, Deep- channel avulsion—Example from the basin floor of the Desoto Can-
water facies, processes and models: a review and classification scheme yon area, Gulf of Mexico (abstract): American Association of Petro-
for modern and ancient sediments: Earth-Science Reviews, v. 23, p. leum Geologists, Annual Meeting, Abstracts Volume, p. A140.
75–174. POSAMENTIER, H.W., MEIZARWIN, WISMAN, P.S., AND PLAWMAN, T., 2000, Deep
PILKEY, O.H., 1988, Basin plains: giant sedimentation events, in Clifton, water depositional systems—Ultra-deep Makassar Strait, Indonesia,
H.E., ed., Sedimentologic Consequences of Convulsive Geological in Weimer, P., Slatt, R.M., Coleman, J., Rosen, N.C., Nelson, H.,
Events: Geological Society of America, Special Paper 229, p. 93–99. Bouma, A.H., Styzen, M.J., and Lawrence, D.T., eds., Deep-Water
PIPER, D.J.W., AND NORMARK, W.R., 1983, Turbidite depositional patterns Reservoirs of the World: Gulf Coast Section SEPM Foundation, 20th
and flow characteristics, Navy submarine fan, California borderland: Annual Research Conference, p. 806–816.
Sedimentology, v. 30, p. 681–694. POSAMENTIER, H.W., AND VAIL, P.R., 1985, Eustatic controls on deposi-
PIPER, D.J., SHOR, A.N., AND HUGHES-CLARK, J.E., 1988, The “1929” Grand tional stratal patterns: SEPM, Research Conference no. 6, Sea Level
Banks earthquake, slump and turbidity current, in Clifton, H.E., ed., Changes—An Integrated Approach, October 20–23, 1985 (Abstract
Sedimentological Consequences of Convulsive Geological Events: and Poster).
Geological Society of America, Special Paper 229, p. 77–92. PRATHER, B.E., BOOTH, J.R., STEFFENS, G.S., AND CRAIG, P.A., 1998, Classifica-
PIPER, D.J.W., PIRMEZ, C., MANLEY, P.L., LONG, D., FLOOD, R.D., NORMARK, tion, lithologic calibration and stratigraphic succession of seismic
W.R., AND SHOWERS, W., 1997, Mass-transport deposits of the Amazon facies of intraslope basins, deep-water Gulf of Mexico: American
Fan, in Flood, R.D., Piper, D.J.W., Klaus, A., and Peterson, L.C., eds., Association of Petroleum Geologists, Bulletin, v. 82, p. 701–728.
Proceedings of the Ocean Drilling Program, Scientific Results, v. 155, READING, H.G., AND RICHARDS, M., 1994, Turbidite system in deep-water
p. 109–146. basin margins classified by grain size and feeder system: American
PIRMEZ, C., AND FLOOD, R.D., 1997, Morphology and structure of Amazon Association of Petroleum Geologists, Bulletin, v. 78, p. 792–822.
Channel, in Flood, R.D., Piper, D.J.W., Klaus, A., and Peterson, L.C., RUST, B.R., 1966, Late Cretaceous paleogeography near Wheeler Gorge,
eds., Proceedings of the Ocean Drilling Program, Scientific Results, v. Ventura County, California: American Association of Petroleum
155, p. 23–45. Geologists, Bulletin, v. 50, p. 1389–1398.
520 HENRY W. POSAMENTIER AND ROGER G. WALKER

SHANMUGAM, G., 1996, High-density turbidity currents: are they sandy Processes of Submarine Fans and Turbidite Systems: New York,
debris flows?: Journal of Sedimentary Research, v. 66, p. 2–10. Springer-Verlag, p. 323–347.
SHANMUGAM, G., LEHTONEN, L.R., STRAUME, T., SYVERTSEN, S.E., HODGKINSON, WEIMER, P., AND LINK, M.H., eds., 1991, Seismic Facies and Sedimentary
R.J., AND SKIBELI, M., 1994, Slump and debris-flow dominated upper Processes of Submarine Fans and Turbidite Systems: New York,
slope facies in the Cretaceous of the Norwegian and northern North Springer-Verlag, 447 p.
Seas (61–67°N): implications for sand distribution: American Asso- WEIMER, P., SLATT, R.M., COLEMAN, J., ROSEN, N.C., NELSON, H., BOUMA,
ciation of Petroleum Geologists, Bulletin, v. 78, p. 910–937. A.H., STYZEN, M.J., AND LAURENCE, D.T., eds., 2000, Deep-Water
SIKKIMA, W., AND WOJCIK, K.M., 2000, 3D visualization of turbidite sys- Reservoirs of the World: SEPM, Gulf Coast Section, 20th Annual Bob
tems, Lower Congo Basin, offshore Angola, in Weimer, P., Slatt, F. Perkins Research Conference, Houston (available only on CD).
R.M., Coleman, J., Rosen, N.C., Nelson, H., Bouma, A.H., Styzen, WILSON J., WALL, G., KLOOSTERMAN, H.J., CONEY, D., CAYLEY, G., WALKER, J.,
M.J., and Lawrence, D.T., eds., Deep-Water Reservoirs of the World: AND LINSKAILL, C., 2005, The discovery of Goldeneye: in Doré, A.G.,
Gulf Coast Section of SEPM Foundation, 20th Annual Research and Vining, B.A., eds., Petroleum Geology: North-West Europe and
Conference, p. 928–939. Global Perspectives: 6th Petroleum Geology Conference, Proceed-
SURLYK, F., AND NOE-NYGAARD, N., 2003, A giant sand injection complex: ings, p. 199–216.
the Upper Jurassic Hareelv Formation of East Greenland: Geologia WINKER, C.D., AND BOOTH, J.R., 2000, Sedimentary dynamics of the salt-
Croatica, v. 56, p. 69–81. dominated continental slope, Gulf of Mexico: integration of obser-
UCHUPI, E., AND AUSTIN, J., 1979, The stratigraphy and structure of the vations from the sea floor, near-surface and deep subsurface, in
Laurentian Cone region: Canadian Journal of Earth Sciences, v. 16, Weimer, P., Slatt, R.M., Coleman, J., Rosen, N.C., Nelson, H., Bouma,
p. 1726–1752. A.H., Styzen, M.J., and Lawrence, D.T., eds., Deep-Water Reservoirs
VAIL, P.R., MITCHUM, R.M., JR., AND THOMPSON, S., III, 1977, Seismic of the World: SEPM, Gulf Coast Section, 20th Annual Bob F. Perkins
stratigraphy and global changes of sea level, part 3: relative changes Research Conference, Houston, p. 1059–1086.
of sea level from coastal onlap, in Payton, C.E., ed., Seismic Stratig-
raphy—Applications to Hydrocarbon Exploration: American Asso-
ciation of Petroleum Geologists, Memoir 26, p. 63–81.
VAN WAGONER, J.C., BEAUBOUEF, R.T., HOYAL, J.C.J.D., DUNN, P.A., ADAIR,
N.L., ABREU, V., LI, D., WELLNER, R.W., AWWILLER, D.N., AND SUN, T.,
2003, Energy dissipation and the fundamental shape of siliciclastic
sedimentary bodies (abstract): American Association of Petroleum
Geologists, Annual Meeting, Abstracts Volume, p. A175.
VAN WEERING, T.C.E., NIELSEN, T., KENYON, N.H., KATJA, A., AND KUIJPERS,
A.H., 1998, Large submarine slides on the NE Faeroe continental
margin, in Stoker, M.S., Evans, D., and Cramp, A., eds., Geological
Processes on Continental Margins: Sedimentation, Mass-Wasting,
and Stability: Geological Society of London, Special Publication 129:
p. 5–27.
WALKER, R.G., 1965, The origin and significance of the internal sedimen-
tary structures of turbidites: Yorkshire Geological Society, Proceed-
ings, v. 35, p. 1–32.
WALKER, R.G., 1967, Turbidite sedimentary structures and their relation-
ship to proximal and distal depositional environments: Journal of
Sedimentary Petrology., v. 37, p. 25–43.
WALKER, R.G., 1973, Mopping up the turbidite mess, in Ginsburg, R.N.,
ed., Evolving Concepts in Sedimentology: Baltimore, The Johns
Hopkins Press, p. 1–37.
WALKER, R.G., 1975a, Generalized facies models for resedimented con-
glomerates of turbidite association: Geological Society of America,
Bulletin, v. 86, p. 737–748.
WALKER, R.G., 1975b, Nested submarine channels at San Clemente,
California: Geological Society of America, Bulletin, v. 86, p. 915–
924.
WALKER, R.G., 1978, Deep-water sandstone facies and ancient submarine
fans: models for exploration and stratigraphic traps: American
Association of Petroleum Geologists, Bulletin, v. 62, p. 932–966.
WALKER, R.G., 1985, Mudstones and thin-bedded turbidites associated
with the Upper Cretaceous Wheeler Gorge conglomerates, Califor-
nia: a possible channel–levee complex: Journal of Sedimentary
Petrology, v. 55, p. 279–290.
WALKER, R.G., 1992, Turbidites and submarine fans, in Walker, R.G., and
James, N.P., eds., Facies Models; Response to Sea Level Change:
Geological Association of Canada, p. 239–263.
WEIMER, P., 1989, Sequence stratigraphy of the Mississippi Fan (Plio-
Pleistocene), Gulf of Mexico: Geo-Marine Letters, v. 9, p. 185–272.
WEIMER, P., 1991, Seismic facies, characteristics and variations in channel
evolution, Mississippi Fan (Plio-Pleistocene), Gulf of Mexico, in
Weimer, P., and Link, M.H., eds., Seismic Facies and Sedimentary
INDEX 521
Index
bioturbation 57, 60, 119, 133–134, 179, 195, 201, 203, 207, 212,
A 249, 251, 253, 258–259, 262, 265, 311–312, 314, 318, 320, 336,
abrasion 23, 137 350, 363, 369, 371, 385, 389, 431
Abu Dhabi 40 Blackhawk Formation 303–304, 321, 333
accommodation 22, 58, 60, 63, 171, 199, 211, 344, 350–351 Bouma sequence 12–13, 369, 399, 403–404, 407
accumulation rate 52, 66, 71 boundary layer 22–23
accumulation surface 32, 50, 52–53, 55, 57, 66–68, 74 bounding surface 13, 15, 19–20, 22, 32, 34, 38–40, 48, 57, 60, 63,
adhesion plane beds 32 67, 73, 159, 179, 453, 478, 485
adhesion ripples 32, 39 braid delta 256
adhesion strata 30, 32, 38 braiding 91, 96–98, 102, 112, 138, 475
adhesion warts 32, 39 Brazil 62, 252, 256, 361, 427–428, 430–432
aerodynamic configuration 53, 55, 60 breaker bars 295, 299–300, 311
Al Liwa sand sea 40 brinkline 31
algal growth 33 brittle failure 33
allocyclic 19, 175, 179, 189, 237, 281, 420 burrow 33, 45–47, 65, 107–108, 122–123, 126, 129–134, 193, 203,
allogenic 50, 60, 67, 73 219, 248–249, 251, 259, 261, 265–267, 280, 314, 329, 363–364,
alluvial architecture 129, 141–145, 147–148, 150, 153–156, 159 385
alluvial fan 24, 50, 71, 80, 85, 119, 135, 137, 139, 140–141, 145, bypass 22, 50, 52, 55, 57, 59–60, 66, 69, 71, 199–200, 216–217,
148, 156–157 246, 267, 280, 344, 359, 365, 401–402, 476, 483, 511
alluvial valley 87, 119, 138, 141, 145, 217, 249
anastomosing 96, 134–135, 138, 139, 160
angle of climb 52–53, 55–56, 62, 66–69, 73, 75–77 C
angle of internal friction 33 Caliente Range 321, 331–332, 337
angle of repose 27, 31–32, 89, 91, 104, 122, 272 California 50, 249, 302, 304, 307, 319–321, 323–325, 328–329, 331–
anhydrite 65 332, 335, 337, 377, 405, 408–412, 414, 429–431, 433, 436, 446–
animal activity 46 447, 454–464
animal trackway (footprint) 33, 43, 46 canyon 8, 113, 279, 304, 361, 363, 373–374, 399–402, 415, 422–
antidune 85–86, 90, 95, 99, 104, 119 432, 454, 485, 493–494, 504, 511–512, 517
arid 19, 22, 24, 70–71, 73, 107, 119, 124, 129–130, 133, 139, 140, capillary fringe 45, 53, 55, 65, 68, 126
156–157 carbonates 157, 200
Arizona 60 Carmel Formation 60, 62
armored lag 26, 58 Carmelo Formation 425
Askja sand sheet 29, 36 CCC turbidite 6, 9, 404, 406, 415, 441, 458–459, 461, 474
Askja, Iceland 39, 44, 50, 55 Cedar Mesa Sandstone 36–38, 43–44, 46–47, 58, 65, 68–70, 77
Atchafalaya 241, 243, 246, 250, 272, 347, 352, 355, 381 cement 22, 55, 62–65, 71, 73, 136, 220, 279, 314, 344, 350, 364, 409
Australia 5, 32, 44, 47, 65, 176–177, 185, 221, 226, 249, 255, 340, cementation 63, 65
342, 360, 366, 369–371, 390–391, 403 channel 1–2, 5–7, 9, 12, 24, 63, 68, 71, 85–87, 91, 93–124, 126,–127,
autocyclic 19, 34, 57, 175, 179, 243, 348, 359, 419–420 129, 133–148, 150–151, 156–160, 171, 176–177, 179–180, 184,
autogenic 73 186, 188–196, 199–200, 202–204, 206–208, 210–211, 216–217,
avalanche 27, 30–32, 40, 104, 107 220, 225, 227, 237, 240–241, 243–244, 246–247, 249–250, 252–
avalanche strata 30–31 253, 257, 260, 262–263, 265–272, 274–275, 279, 281, 285, 300,
avulsion 119, 121–122, 134–143, 147, 150–151, 156–160, 237, 243, 312–313, 316–317, 331, 353, 361, 367–369, 372, 399–404, 412,
249, 252–253, 266–267, 274, 285, 421, 441, 448, 463, 501 414–451, 453–454, 456–459, 461, 463–465, 468–470, 472–478,
482–485, 486, 489–490, 493, 495–496, 501–502, 504, 510–511,
513–514, 516
B channel belt 85, 87, 91, 97, 102, 108–112, 115–122, 126, 134–139,
Bahrain 50 141–144, 146, 151, 156, 158, 270, 438, 449, 476, 484
bar and trough system 316 channel fill 2, 6, 91, 96, 100–101, 104, 107–110, 112, 114, 117–118,
barchan (barchanoid) 28, 32, 34, 40–43, 53, 63 122, 133, 141, 159, 257, 265, 401, 425, 427–432, 436, 440, 446,
barrier island 10, 200, 240, 244, 260, 271, 297–298, 300, 303, 381, 389 449, 457–458, 461, 463, 475–476, 482–484, 504
basalt 25, 62–64 channel migration 86, 96–97, 99, 103, 119, 137, 141, 176, 190, 253,
baseline of erosion 58 266, 434
bedform 19–20, 22, 27–30, 32, 34–40, 43, 45–53, 55–57, 62–64, 66, channel pattern 91, 93, 96–99, 102–103, 116, 118–119, 134–135,
68, 70, 72–73, 76, 85, 89–91, 102, 104–105, 113, 174, 221, 243, 137–138, 140, 143–144, 147–148, 156–159, 186, 203, 253, 419,
270–271, 295, 306–307, 311, 313, 361, 363, 366, 371–372, 385, 425, 427
403–404 chenier 243, 268, 271–272, 355
bedform behavior 34 climate 19, 22, 43, 46–47, 58, 60, 63, 69, 71, 73, 80, 85, 107, 119,
bedform climb 22, 43, 50, 52, 56–57, 62 121–122, 124, 126, 130, 133–135, 139, 141, 143, 147–149, 151,
bedform migration 19, 22, 34, 39, 48–49, 52, 56, 66, 68, 73 155–160, 171, 178–179, 199, 216, 225, 241, 254, 307
bedload 23, 86–87, 89, 94, 97–99, 104, 119, 137, 182–183, 186, 191, climate change 22, 58, 60, 69, 71, 73, 122, 135, 141, 143, 148–149,
243, 267, 295, 353, 361 151, 156–158, 241
bedload sheet 86–87, 89, 98–99, 104 climatic cyclicity 71

Facies Models Revisited


SEPM Special Publication No. 84, Copyright © 2006
SEPM (Society for Sedimentary Geology), ISBN 1-56576-121-9, p. 521–527.
522 INDEX

climb 3, 36, 40, 52, 53, 55–56, 62, 66–69, 75–76 Devonian 5, 11, 44, 63, 112, 114–115, 123, 129, 131, 157, 238, 253–
clinoform 237, 239, 250, 262, 271–276, 280, 362, 363, 367, 385 254, 274, 403, 406
coastal 1, 22, 27, 43, 45–47, 50, 57–58, 60, 62, 73, 119, 139, 147, 150, Dhahran 45, 50
157–159, 171, 175–177, 179–182, 191–193, 197, 215, 217–218, diachronous 58, 199
221, 226–227, 237, 240, 249, 254, 258, 293, 297–298, 300, 303, diagenetic cement 55
305–306, 318–320, 323–324, 327,332, 344, 347, 349, 350, 357, diastem 58, 191, 194, 390
367, 461, 493 directional variability 27
coastal classification 180, 182, 218, 221, 226 distributary channel 9, 199, 202–203, 237, 241, 243–244, 246, 249–
cohesive 22, 32, 44, 97, 425 250, 252–253, 257, 260, 262–263, 265–272, 274–275, 279, 417,
cold-climate desert 47 422, 427, 431, 463, 468, 473
Colorado 11, 40, 60, 62, 201–202, 210, 373, 391 distributary mouth bar 3, 9, 237, 241, 246, 250, 253, 263, 344
Colorado Plateau 60, 62 diurnal 27, 34, 156, 361, 364
compaction 58, 142, 143, 158, 404, 449 draa 27–31, 35, 43, 51, 62–63
compound bar 86, 91, 94, 98, 100–102, 104–105, 108 drift potential 28, 34
contorted bedding 46 drift potential (resultant, RDP) 28, 34
core 3, 5, 9, 12, 39, 50, 70, 73, 116, 201, 203, 218, 255, 272, 275–277, drift direction (resultant, RDD) 28, 34
304–305, 319, 363, 367, 379, 381–382, 386, 407, 427, 431, 452 dry eolian system 19, 52–53, 55-56, 60–62, 68, 74
crawling trace 33, 129 dry interdune 31, 43, 45, 66, 68, 78
creep 23, 26–27, 37 Duero Basin 46
Cretaceous 3–4, 6, 9–12, 36, 62–63, 153, 157, 174, 177, 180, 184, dune 13, 19–20, 22–23, 25, 27–36, 38–48, 50–57, 60, 62–63, 65–76,
186, 191–192, 196–197, 199–200, 202, 204, 206–208, 237, 243, 78, 85–91, 94–95, 98–100, 104, 107, 109, 112, 115–116, 119, 121–
248, 253, 261–263, 266–268, 270–271, 273–274, 276–281, 283, 122, 139, 187, 191, 227, 268, 295, 312, 316–317, 319, 329, 361,
285, 302–303, 333, 341, 374–376, 405, 407, 410–411, 414, 427, 365–366, 370, 372
430, 446–447, 454, 501 dune element 40, 43, 45–46, 54, 62–63, 66–67, 69
crevasse splay 91, 115, 119–124, 126–127, 134–135, 137, 139, 270, dune flank 27
400, 422, 438, 463, 468–474, 504 Dunvegan Formation 180, 191, 248, 261, 266, 268–270, 274, 276–
crinkly laminae 47 282
critical climbing 52 dynamic facies model 19, 22, 67, 69, 73, 77
cross bedding (cross-bedding) 2, 6–8, 10–11, 19, 32, 34, 38–39, 42,
190, 249, 258, 261, 265, 298, 307, 311–313, 317, 319, 322–323,
331, 363–364, 366, 371, 373, 375, 380, 389, 407 E
cross strata 32, 33–35, 38–41, 48–50, 52, 55, 63, 86, 89–90, 94, 98– Ebro delta 252
100, 104, 107, 109, 112, 121–122, 127, 268 ejecta 23
cross stratification 7, 13, 15, 32, 78, 88, 104, 195, 249, 258, 260, 281,
England 3, 10, 13, 43–44, 47, 62, 65–66, 191, 215, 364, 413
381 Entrada Sandstone 48, 62, 65
crust 33, 46, 129, 134, 148, 149, 153, 157, 353 entrainment 19, 22–23, 25, 50, 53, 119, 351
curl 33, 45, 46 eolian 6, 8, 13, 19–20, 22, 23, 25–40, 43–48, 50, 52–58, 60–63, 65–
current ripple(s) 33, 107, 126, 258, 260, 375, 427, 442, 458, 468, 473 76, 78, 80, 139, 153, 157, 270, 272, 329, 415
eolian plane beds 43
eolian system, wet 19, 53, 55–57, 61–62, 68–69, 73, 75
D eolian system, wet–dry 74
damp interdune 32, 38, 45, 47, 53, 65, 68 ephemeral 46, 50, 62, 119, 122, 126, 129, 133, 139, 153, 156, 176,
Danube 240, 244, 249, 255, 259–260, 267, 270 180, 237
deflation 19, 22, 26, 28, 43, 47–48, 52, 55, 57–61, 65–66, 69–70 erg 22, 24, 45–46, 50, 53, 57–58, 60, 62– 63, 65, 66–67, 69–71, 75–
delta 2–3, 6, 9–10, 13, 16, 50, 85, 107, 115, 119, 121–122, 134–135, 77, 80, 88, 143–145, 176, 180, 239, 257, 272, 297
138–139, 143, 153, 157–158, 173–174, 176, 178–180, 182, 184– erg construction 22, 50
185, 189–191, 194, 196, 199–204, 206–208, 210, 215–220, 225, extra-erg 46
237–268, 270–282, 284–285, 293, 320, 324, 339, 342, 344–345, estuarine lithosome 189, 190
347, 349–350, 352–353, 355, 360, 363, 367, 371–373, 375, 381– estuary 5, 171, 173, 176–177, 180, 182–191, 193, 195, 198–201, 206–
383, 390, 401, 425, 427–429, 431, 436, 463, 471, 493 208, 212, 215–221, 225–226, 237, 256, 265, 268, 282, 285, 344,
delta front 184, 246–247, 249–251, 253, 256, 257, 259, 261–262, 363, 372, 401–402, 445
266–268, 270–272, 279 Etendeka igneous province 62
delta, lacustrine 115, 119, 121–122, 139, 249, 276 Etjo Formation 36, 62–63
delta plain 157–158, 184, 237, 244, 249–250, 253–257, 265, 270, eustasy 147, 158, 254, 281, 297, 339
272, 355, 381, 431, 493 evaporite 33, 47, 65, 124, 129, 156–157
density 22–23, 37, 89, 129, 153, 179, 203, 241, 243, 246, 258, 347, evaporite precipitation 33, 47
369, 370, 401–403, 408, 415–416, 421, 425, 427, 493, 504
density current 369, 401
depositional elements 2, 9–11, 13, 15–16, 24, 183, 399–401, 403, F
414–415, 420, 422, 428, 431, 447, 470, 484–485, 501, 504 facies architecture 19, 69, 172, 237–238, 240, 255, 258, 264, 268,
depositional model 64, 66–68, 85, 86, 99–100, 111, 119, 121–124, 270, 281, 388
135, 171, 293, 355, 414 facies association(s) 1, 5–7, 13, 15, 54, 71, 298–299
desert 13, 19, 22–23, 33, 39–40, 45–47, 50–51, 58, 63, 66, 129 facies definition 2
desert rose 33 facies model(s) 1–3, 6–10, 12–13, 15–16, 19, 22, 39, 66–67, 69–70,
desiccation cracks 11, 33, 45–46, 57, 107, 119, 122, 126, 133, 156 72, 73, 75, 77, 85, 171–172, 174–180, 182–183, 189, 191, 212,
INDEX 523

215–218, 220, 223, 225, 237, 282, 293–298, 300, 302, 304–305, Great Sand Dunes 40
311, 322, 324, 339, 341, 345, 347, 363, 369, 379–380, 390, 399, ground-penetrating radar 22, 85, 179, 270
430, 504 growth fault 243, 256–259, 263, 274, 279
facies modeling 1–3, 13, 16, 66–67, 177, 430 Guerrero Negro 46, 58
facies succession(s) 1–2, 6–7, 71, 177, 179, 183, 215, 237, 240, 255, Gulf of Mexico 9, 147, 200, 237, 239–240, 243, 253, 272, 275, 305,
257–258, 260–261, 264, 267, 272, 277–281, 304, 504 339, 341, 348–349, 352, 356, 369, 372, 378, 383, 390, 423, 426–
fair weather (fair-weather) 9, 15, 250–251, 294, 296, 300, 311–314, 427, 429, 431, 437, 439, 441–443, 446–447, 449, 453–454, 469–
317–320, 323–324, 326, 328, 336, 356, 360, 369, 377 470, 475–476, 478, 485–486, 489, 492, 494, 498, 501–502
fenestral porosity 33 gutter casts 317, 323, 367, 374
Ferron Sandstone 262–263, 268, 271, 283 gypsum 33, 46, 50, 65, 129
flake 33, 46
flash flood 46
flood basalt 62, 64 H
flood basin 115, 117, 119–121–122, 126, 134, 136–137 halite 33
flood deposit 107, 137, 255–256, 259, 377 hard-pan 46
flood plain 85–87, 96, 110, 113, 115, 117–123, 126, 129–130, 134– Helsby Sandstone 43–44, 62, 66–68
139, 141–145, 147, 150, 153, 156–157, 159–160, 179, 186, 192, high-energy coast 293, 304, 311–314, 319–320, 324
194, 196, 201, 210, 237, 255, 267 Holocene 14, 43, 113, 116, 134, 143, 157, 217, 247, 293, 302, 339,
flow concentration 52 343, 349, 351, 367, 379
flow fluctuation 34 homopycnal 241, 249
flow reattachment 27 humid 22, 43, 58, 107, 121, 124, 139, 156
flow separation 27, 93, 95 hummocky cross-stratification 301, 317, 370–371, 373–374
fluid mud 243, 246, 253, 311, 345, 355 hyperpycnal 241, 246–247, 249, 253, 258, 282, 344, 347–348, 350,
fluvial 2, 6, 9–11, 15, 19, 27, 43, 46, 50, 62–63, 65–66, 68–71, 85–88, 353, 356, 358, 370, 374, 402, 415, 427, 493
111, 113, 129, 131–135, 141–144, 146–150, 153, 156–160, 171– hypopycnal 241, 243, 249,–250, 258, 344–345, 348, 352–353, 355,
172, 174–176, 178–180, 182–183, 186, 189–192, 194, 195–202, 358
204–205, 207, 209–212, 215–218, 225–226, 237, 240, 242, 249,
253, 255–256, 258, 260–262, 265, 270, 272, 277, 279–282, 331,
339, 344, 347–348, 350, 353–354, 359, 415, 425, 427, 434–435, I
444 Iceland 29, 36, 39, 44, 47, 50, 55
fluvial inundation 46 ichnofacies 129–130, 193, 195, 219–221, 251, 259, 265, 280–281,
fog 65 300, 302, 305, 369, 373, 375, 377, 381–382, 385–386
forced regression 257, 260, 279, 293, 299–301, 318, 320–322, 324, ichnology 171, 195, 218, 281, 379, 385, 386, 390
349, 354, 431, 436, 493 impact zone 27
forcing parameter 34, 71 incised valley 145, 159–160, 171–175, 177, 179–180, 184, 191–193,
foreset 10, 13, 30, 39, 41, 43, 47, 49–50, 238–239, 249–250, 272–274, 195–201, 206, 208, 215–217, 219–220, 223, 253–254, 279–280,
320, 329–331 341, 372, 374, 389, 493, 511
foreset azimuth 39, 43, 49 inclined strata 86–87, 99–105, 107–108, 110–111, 113–114
fossil 4–5, 15, 47, 123, 129–134, 157, 182, 190, 203, 207, 209, 212, Indian Ocean 360
215, 218–219, 249, 300, 304, 359, 379 infiltration capacity 46
friction 22–23, 33, 112, 183, 186, 241, 243, 246, 249, 253, 258, 267, interdistributary bay 3, 243, 249, 253, 268, 270
270, 350, 356, 360 interdune 19–20, 22, 27, 31–33, 38, 40, 43, 45–48, 51–57, 60, 62–63,
frontal splay 250, 270–271, 399–400, 415–417, 419–422, 447, 473– 65–71, 73–76, 78
478, 484–485, 493, 504, 508, 511 interdune flat 38, 43, 45, 52–53, 55, 57, 62, 68–69
Frontier Formation 261, 273–274, 277, 279, 284 interdune hollow 38, 51, 56, 66–68
interdune migration surface 38, 48, 56
interglacial 43, 156–158, 201, 346
G intra-erg 46, 63
Gallup Sandstone 265 inverse grading 27, 30, 37, 247, 403, 407
gas 1, 3, 19, 70–71, 79, 85, 123, 129, 133–134, 157, 159, 202, 210, isochronous 58
221, 237–238, 244, 266, 278, 341, 362, 383, 407 isolated bedforms 45
geomorphic element 22, 40, 47 isolated sandstone bodies 299
Gilbert delta 250
glacial 19, 22, 43, 50, 97, 135, 139, 156–158, 201, 246, 293, 302, 346,
366 J
Gondwana 62 Jackson Group 320, 328, 330
graded bedding 258, 399, 404, 407–408 Jafurah area 46
grain packing 32 Jurassic 48, 60, 62, 65, 133, 153, 354
grainfall 30–33, 36, 38, 40–41, 78
grainfall strata 30–32, 40
grainflow 27, 31–32, 37, 40–41, 43–44, 78 K
grainflow, amalgamated 31 Kelso Dune Field 50
grainflow strata 31–32, 44, 78 Kuiseb River 27
Gran Desertio 40 Kuwait 50
524 INDEX

Mekong 243, 251, 253, 261, 270


L Mexico 9, 19, 40, 45–46, 50, 62, 147, 157, 200, 237, 239–240, 243,
lacustrine 27, 50, 107, 115, 119, 121–122, 124, 127, 129–130, 139, 253, 265, 272, 274v275, 298, 305, 309, 339, 341, 348–349, 352–
146, 156, 180, 200, 238, 240, 249, 268, 276 353, 356, 369, 372, 378, 383, 390, 423, 426, 427, 429, 431, 437,
lacustrine delta (see delta, lacustrine) 439, 441, 442, 443, 446, 447, 449, 453–454, 469, 470, 475–476,
lag 2, 7, 9–15, 26, 28, 30, 43, 47, 50, 58, 63, 91, 148–149, 172–173, 478, 485–486, 489, 492, 494, 498, 501–502
180, 182, 193, 201–204, 217, 221, 237, 239–240, 249, 256, 260, migration 19–20, 22, 27–28, 30, 32, 34–35, 37–40, 43, 48–49, 52,
264, 267, 270, 275, 280, 284, 297, 314, 320, 322, 349, 353, 365, 55–56, 58, 62–63, 66, 68, 70, 73–75, 85, 86, 89–91, 96–97, 99,
368–369, 402, 408, 430, 432, 493 101–105, 109–112, 115, 119, 121–122, 124, 137, 141, 143, 148,
Lagniappe delta 275 159, 176, 179–180, 184, 190–191, 210, 243, 253, 266, 268, 272,
lagoon 7, 9–10, 13–15, 173, 180, 182, 215, 217, 237, 239–240, 242, 279, 297, 300, 306–307, 313, 362–363, 372, 379–380, 383, 385,
249, 256, 260, 264, 267, 270, 493 387, 390, 424, 428, 434, 435, 439, 441–444, 446, 451–452, 475,
lagoonal 10, 13–15, 237, 240, 264, 267 483
lake 24, 32–33, 45–47, 50, 57–58, 60, 62, 71, 80, 102, 113, 119, 121– migration direction 28, 34, 39–40, 48, 49
124, 127, 129–130, 134–135, 137–139, 143, 153, 156–157, 180, migration speed 34
184–185, 191, 197, 201–202, 206, 238–239, 246, 249, 253, 414 Milankovitch 58, 62, 157, 158
Lake Lucero 50 mineralogy 22, 157, 379, 388
laminar flow 25 Mississippi 8–9, 96–97, 116, 121, 157–159, 174, 176, 180, 191, 197,
lateral accretion 434–435, 441, 443–446, 449, 483 201, 237–238, 240, 242–243, 246, 249–250, 253–254, 258, 260,
lee 20, 27–28, 31–35, 37–39, 48, 52–53, 91, 104, 216, 305, 311, 315, 264, 270, 272, 285, 344, 347, 349, 352, 355, 371, 373, 381–382,
363 390, 401, 422–427, 471, 512
lee-side depression 27, 31, 38 Mississippian 180, 191, 197, 201
Leman Sandstone 62, 78 mixed eolian system 65, 69
Lena River 252 Mobile Bay 215, 352
levee 1–2, 6, 9, 12, 91, 115, 119–124, 126, 134, 137, 139, 179, 189, moisture 22, 43, 45, 52–53, 65, 126, 130
199, 243, 270, 399–401, 404, 406, 412, 414–422, 425, 427, 429, mojave 22, 43, 45, 52–53, 65, 126, 130
431–432, 437–443, 446–447, 449–451, 453, 456, 458–459, 461, morphology 19, 27, 38–39, 42, 45, 50, 55, 62, 66–67, 70, 73, 171,
463–464, 468–470, 472–478, 485, 490, 493, 494, 496, 501–502, 174–176, 178, 182–184, 215, 227, 237–238, 240–242, 253, 255–
504, 510–511, 513–514 256, 316, 323, 339, 390, 417, 419, 422, 428, 471–472, 477, 501,
leveed channel 1, 400, 401, 415–417, 419–422, 427, 429, 431, 437– 504
438, 441–443, 447, 450, 473–478, 485, 493, 496, 504, 511 mud curl 45–46
liquefaction 32–33, 43 mud drape 46, 50, 107, 195, 198, 203, 205, 262, 267, 361, 373, 375
lobe 1, 2, 91, 96–97, 104, 119, 139–141, 158, 237, 240–242, 244, 246, mud flake 33, 46
250, 255–256, 260, 262, 264, 267–268, 270–271, 274–275, 277– muddy substrate 46
281, 340, 343, 381, 390, 399–401, 415–418, 420, 422, 473, 475–
476, 484, 496, 501
loess 23 N
low accommodation 175, 191, 199, 202, 209 Namib Desert 45, 47, 50–51, 63
low-energy coast 317–319, 324, 333 Namib Sand Sea 27, 32, 35, 45–46, 50, 52
Lower Cutler Beds 36, 45, 63, 68 Namibia 27, 29, 32, 36–37, 45–46, 55, 58, 62–63
lower shoreface 300–301, 318, 320–323, 329, 340 Navajo Sandstone 35, 62, 68
lower-stage plane bed 89, 95, 104 nearshore circulation cells 295, 299, 311, 316, 321
lunate megaripples 307, 312, 316–317, 319, 323, 327 nearshore profile 313, 320
Nebraska Sand Hills 43, 46
negative sediment budget (see sediment budget negative)
M nested reactivation surface 34
Macaronichnus 300 net sediment budget (see sediment budget net)
Mahakam 250 neutral sediment budget (see sediment budget, neutral)
marine flooding 57 New Mexico 19, 45, 50, 62, 265, 272, 353, 372
marine regression 60, 149, 159, 297 Niger 47, 261
marine, shallow 16, 19, 50, 159, 176, 246, 280, 293, 297, 302, 308, Nile River 174
320, 324, 349, 386 Nile delta 238, 249, 270
marine transgression 58, 62, 149, 156, 159–160, 199, 297, 300, 323 North Sea 19, 58, 62, 70, 78–81, 358, 375
mass-transport complex 401, 412–413, 485–491, 493, 498–499, 504, numerical modeling 220–221
508
Mauritania 43
meandering 7, 9, 12, 15, 91, 96–98, 101–102, 108, 112, 116–117, O
135, 138, 140, 142, 156, 175, 186, 188, 190–191, 196, 199–202, oblique migration 35, 63
209–210, 212, 268, 401, 427, 434–435, 444, 501 Oman 25
meandering channel 97, 101, 188, 401, 434–435, 501 Ophiomorpha 265, 300, 382
mechanical loading 33 optically stimulated luminescence 22, 58, 85
mega-bedforms 30 orbital forcing 58, 158
megaripple 27, 29, 47, 306–307, 311– 312, 315–317, 319, 323, 327, Oregon coast 45, 312, 320
370 Ormskirk 62, 68
INDEX 525

overbank 71, 94, 107, 115, 119, 121, 124, 136–138, 140, 143, 156, ravinement 189–191, 194, 199–201, 206, 209, 218, 224, 251, 279–
158, 160, 194, 199–200, 210–211, 415, 417, 422, 438, 441–442, 280, 297, 300, 349, 351, 372, 376, 381–382, 386, 388
450, 452–453, 463, 468, 473–474, 476, 482–484, 501, 504 reactivation surface 34–35, 38, 48, 65, 195, 261–262, 375
recognition criteria 197, 217, 361
reg 26
P regional climate 46, 58, 157
Padre Island 41, 45–46, 305, 308, 320, 326 regression 15, 60, 113, 149, 159, 180–181, 192–193, 197, 215, 257,
Page Sandstone 60 260, 277, 279, 293, 296, 297, 299–301, 318, 320–322, 324, 337,
paleoenvironment 19, 57 349, 354, 431, 436, 493
paleosol 33, 46, 68, 113, 126, 128–130, 156–160, 193–194, 196, 201– remote sensing 85
202, 209–211, 267 reptation 23, 26–27
parabolic 33, 40 reservoir prediction 70, 73
Paradox Basin 63 resultant drift direction (see drift direction, resultant)
Paraíba do Sul 252, 259, 271 resultant drift potential (see drift potential, resultant)
Paraná Basin 62 reversing dune 40
parasequence 196–197, 202, 268, 279, 294, 297, 300, 321, 383, 386, rhizoliths 33, 47, 57–58, 60, 69, 133–134
388 Rhone (Rhône) (River, delta, fan) 8, 259–260, 274, 401, 465, 501
particle size 22, 25 rip channel 312, 316–317
pebble lag 7, 43, 193, 201–202 rip current(s) 178, 293, 295, 311, 313–314, 317, 320–321, 357, 366
Pennsylvanian 3, 36, 63, 68, 180, 253–254, 272, 274 ripple 3–4, 6, 9, 11–13, 27–33, 36–41, 43, 45–47, 52, 56, 78, 85–87,
perched water table 46 89–90, 94–95, 98–100, 104, 107, 112, 119, 122–123, 126–127,
permafrost 63 191, 203, 207, 210, 248–249, 258, 260–262, 298, 301, 306–307,
permeability 46, 85, 107–108, 110, 144, 201–205, 207, 210, 211– 311–320, 323, 327, 331, 363–365, 369–370, 373–375, 399, 403–
212, 404 406, 427, 442, 446–447, 458–459, 467–468, 473, 481
Permian 8, 19, 36–38, 43–47, 58, 60, 62–63, 65, 68–71, 77–81, 180 ripple indices 27
pinstripe lamination 31, 36 ripple strata 30–31, 36–38, 41, 43, 45, 78
planar strata 86, 89, 99, 104, 122 ripple trough 27, 30, 37
plane beds 32, 43, 47, 86, 89, 95, 99, 104, 119, 122 rippleform laminae 30
plant colonization 46, 58 river delta 252, 273, 367, 471
plant root(s) 33, 63, 121, 126, 129, 133–134, 156 river mouth 176–177, 189, 199, 215–216, 218, 226, 240, 249, 253,
playa 45, 47, 71, 139 255, 257–258, 267, 282, 304, 344–345, 347, 350, 356, 402, 432,
Pleistocene 4, 14, 43, 113, 157–158, 198, 223, 238, 320, 323–324, 493, 511
328, 329, 335, 349, 375, 379, 402, 429, 437, 449 river-influenced 9, 240
Pleistocene terrace deposits 328–329 root structure 33, 45, 47, 63, 330
plinth 27, 32, 36 Ross Sandstone 434, 445, 474, 476, 479–480, 483–484
Po delta (Po River) 243, 247, 252 Rotliegend 19, 58, 70–71, 73, 78, 80–81
point bar 1, 12, 91, 94, 96, 98–102, 104, 109, 118, 186, 212, 434–
435, 445
polygon 33, 57, 60 S
polygonal fracture 57, 60 sabkha 24, 47, 50, 55, 60, 62, 71–72, 80
pore-water pressure 32 Sahara 47, 52
potential sediment load 50 Salima sand sheet 47
pour-in texture 30 salt crust 33
primary airflow 27, 55 salt flat 24, 47
prodelta 184, 190, 199, 218, 237, 239, 248–249, 251, 253, 255, 257– salt-growth 47
258, 260, 262–264, 267–268, 270–272, 274, 280, 347 salt precipitation 47
progradational shoreline 297, 299 saltation 23, 27, 30, 37–38
prograding strand plain 298 saltation cloud 23
Proterozoic 60, 65, 186 saltation path length 27, 37
pseudomorph 33, 46, 65 San Clemente 429–431, 433–436, 501
pumice 25 San Miguel Formation 280, 285
sand sea 20, 22, 27, 32, 35, 40, 45–46, 50, 52, 40
sand transport direction 27–28, 33
Q sand-accumulating bedform 28
Quaternary 22, 43, 58, 73, 134, 156–158, 180, 199, 210, 239, 272, sand-transporting bedform 28, 34
339, 341, 343, 348, 354, 359, 363–364, 371–374, 375, 381, 385, sandflow 31, 34
388, 389, 390, 402 sandsheet 29
Queensland 65, 371 satellite 22, 24, 85, 221, 340–342, 352, 369, 391
Saudi Arabia 45–46, 50
scarp recession grainflow 31
R scour pit 39, 48
raindrop imprint 33, 122 sea level 2, 7, 9–10, 13, 22, 50, 58, 73, 122, 124, 141, 148, 150, 152–
rainfall 30–33, 36, 38, 40–41, 46, 78, 130, 149 153, 156–160, 171, 173–175, 179–182, 184–185, 189–192, 197–
526 INDEX

202, 215–218, 220, 224–225, 227, 238–239, 246, 254–256, 279, slab slide 33, 37, 44
281–282, 285, 293–294, 296–297, 299–303, 306, 324, 337, 339, slipface 27, 28, 31–32, 40, 62–63, 127
346–349, 354, 359, 390, 399, 401–402, 425, 431, 436, 490, 493, slope channel 399–400, 422, 425, 427–429, 431, 437, 441, 444, 451,
501, 511 490, 504
seabed imagery 221 slump degradation grainflow 31
seasonal 27, 34, 46, 68, 119, 121, 129–130, 133, 156, 241, 249, 360, slumping 32, 412, 425–426, 441, 466, 474
363 soft-sediment deformation 32, 43–44, 258, 260, 274, 458–459, 461
secondary airflow 27, 35 soil 119, 126, 128–130, 133, 138, 156, 160, 193, 353
sediment availability 19, 50, 68–70, 73, 76–77, 121, 148 sole marks 367, 374
sediment budget, negative 47, 57 South Atlantic 62
sediment budget, net 40, 52, 55, 57, 59, 61, 67 southern North Sea 19, 58, 62, 70, 78–81, 358, 375
sediment budget, neutral 57 Spain 45–46, 252, 319–320, 324, 327–329, 335
sediment compaction 58 spatial complexity 73
sediment flux 22, 57, 67, 149, 179–180, 182–183, 218, 220, 504 sphericity 25
sediment supply 2, 9–10, 13, 19, 22, 27, 30, 50, 55, 64–65, 69, 85, staging area 399–400, 415, 417, 422, 425, 485, 491, 493, 504
91, 122–124, 135, 137, 138, 141, 144–145, 148–149, 151, 156– static facies model 67, 174
159, 171, 179–182, 191–192, 197, 199–200, 216, 225, 250–251, stepped transgression 297, 300, 321
255, 258, 279, 281, 297, 303, 341, 344, 347, 349–350, 355–356, Stokes surface 19, 57, 59
374, 380, 390, 422, 501 storms 14, 86, 90, 139, 241, 247, 249, 253, 258, 293, 295, 313–314,
sediment transport 23, 50, 52, 85–86, 91–95, 98–100, 107, 112, 119, 317–319, 339, 350, 356–357, 359–360, 366–367, 369–370
135–139, 141, 145, 147–149, 158, 176, 179, 186, 217, 221, 242, stoss 27, 29, 36–37, 52, 91, 305
258, 295, 320, 343, 350–352, 360, 362, 366, 369, 370, 511 strataset 85–87, 90–91, 102, 108, 110, 115, 121–123, 137
sediment transport rate 50, 52, 91–93, 95, 99, 135,–136, 139, 141, stratigraphic modeling 22
145, 148–149, 179, 242 stratigraphic organization 175, 198, 200, 281
sediment waves 400, 422, 437, 446, 450–451, 453–454, 463, 468, subsidence 11, 46, 57–58, 61, 65, 67, 119, 135–136, 138–139, 141–
472, 475, 504 142, 147–151, 153, 156, 159, 237, 257, 274, 279, 282, 297, 302,
sedimentary basin 137, 141, 147–149, 180, 202, 211, 219, 220 339, 342, 381
seif 28, 144, 146 superimposed bedform 20, 30, 35, 39, 40, 48, 102, 104–105, 366,
seismic 1–2, 5, 70, 85, 115–116, 118, 141, 144–145, 159, 171, 175, 385
179, 192, 197, 203, 206, 216, 218, 220–221, 223, 238, 239, 256– superimposition 27, 35, 38, 40, 43, 48, 62
257, 267, 270, 272, 274–275, 305, 362–364, 367–368, 372, 375– supersurface 22, 48, 55, 57–60, 62, 66, 67, 69–71
376, 383–385, 399, 401, 413, 415, 417, 422–426, 428–429, 431, supratidal flat 65
434, 437–444, 446–447, 458, 468–470, 473–477, 484–486, 489, surface creep 23, 26
490, 496–498, 500–501, 503–508, 510, 512 surface stabilization 22, 55, 58
semiarid 22, 70, 119 surface trace 45
separation cell 27 suspended load 23, 97, 119, 255, 311, 345, 353
sequence 11, 13, 22, 27, 32, 40, 48, 50, 52, 57–58, 60, 66–69, 71, 73, swaly cross-stratification 301, 311
77, 91–92, 100–101, 103–104, 107, 115–116, 122, 124, 126–127, systems tracts 193, 199, 216–218, 238, 281–282, 339, 348–349, 359,
129–130, 134, 137, 141–145, 156, 157, 158–160, 171, 175, 179, 386, 388, 390, 425
189, 191–200, 202, 209, 211, 215–216, 218–219, 221, 237–239,
241, 253, 255–256, 260, 266, 268, 274, 279, 281–283, 294–295,
297, 299–300, 318, 320–322, 325, 328–339, 341, 348, 351, 369, T
376, 383, 385–389, 399–401, 403–404, 407, 414, 420, 431, 457, tectonism 58, 85–86, 119, 122, 124, 135–138, 141, 143–144, 151,
459–463, 475, 483, 485, 493, 510 156, 158–160, 293
sequence boundary 57, 189, 191–194, 197–200, 202, 209, 215–216, teepee 33, 47
218, 348, 388–389 Ténéré 47
sequence stratigraphy 159, 171, 175, 239, 274, 281–282, 299, 341, Tertiary 63, 132, 134, 148, 153, 157, 359, 383, 431, 458, 501
385–386, 401, 485 Texas 41, 45–46, 180, 200–201, 238, 249, 259, 281–282, 285, 308, 310,
serir 26 312, 320, 324, 326, 328, 330, 336, 349, 353, 369,–371, 373, 378
setdown 295, 299, 321 tide 3, 8–10, 171–172, 176–179, 181–184, 186, 189–191, 196, 197,
setup 295, 299, 321, 350, 357, 366–367 215, 217–218, 225, 237–241, 243, 249–256, 261–262, 266–267,
shadow zone 27, 37 270, 273, 277, 279–280, 282, 285, 293, 307, 344, 349–350, 359–
shallow marine (see marine, shallow) 364, 366, 372, 375, 380
shear stress 23, 25, 88–89, 91–95, 112, 119, 137, 139, 243, 350, 360, tide-dominated 8, 171, 177–179, 183–184, 186, 189–191, 196, 215,
368 217–218, 237, 240–241, 252–254, 256, 267, 270, 277, 282, 344,
Shikaoda Formation 65 362–363, 372
shoreface 180, 184, 189–190, 196, 199–200, 202, 207–208, 210, 215, tide-influenced 239, 241, 251, 254, 256, 261–262, 266, 270, 273,
217–218, 221, 224, 237, 249–251, 258–260, 265, 267, 271–272, 279–280
278–279, 281–282, 293–305, 307–314, 317–324, 327, 329–333, top-truncated delta 251, 279, 281
339–341, 349–350, 355, 357, 366–367, 370–374, 380–381, 385, topset 238–239, 249, 251, 260, 272, 279, 281
390 trackway 33, 43, 46–47, 134
shoreface profile 299–300, 304, 308–309, 323, 327, 339 transgression (transgressive) 10, 14–15, 58, 60, 62, 149, 156, 159–
shoreface-attached ridges 305, 310, 370–372 160, 171–172, 176, 179–182, 184, 189–192, 194–202, 207, 210,
Silurian 63, 157, 179 215–218, 224, 237–238, 240, 251, 255–256, 260, 265, 277–278,
Skeleton Coast 27, 29, 32, 46, 55 280–282, 284–285, 293–294, 297, 299–300, 303–305, 308, 321,
INDEX 527

323, 332, 339, 343–344, 348–349, 351, 354, 359, 362–364, 367,
371, 374, 379, 380–381, 383, 386–388, 390, 425, 485, 493 W
transition point 417–418, 420, 473, 475–476, 485, 493, 508 Waitemata Group 415, 461, 465–466, 468
translatent rippleform stratification 30 Walther’s Law 237, 293
transport 2, 19, 22–23, 25–28, 30–31, 33–35, 38, 40, 45, 47, 50–53, water flow 86, 91, 93–94, 99, 107, 119, 137, 312
55, 57, 63, 67, 85–86, 89, 91–95, 97–100, 107, 112, 119, 122, water table 19, 22, 33, 45–47, 50, 52, 53, 55–63, 65–70, 74–75, 77,
129, 133–139, 141, 145, 147–149, 158–59, 171, 176, 179–180, 85, 126, 130, 133, 138, 160, 193
183, 186, 196, 200, 217, 221, 225, 242–243, 253, 258, 295, 298– water-table elevation 46
299, 303, 306, 311, 314–315, 320, 343–345, 348, 350–353, 356, wave base 15, 294–296, 318, 328, 369, 399, 493, 511
359–363, 365–366, 369–371, 373–374, 377, 390, 399, 401, 407– wave energy 183–184, 250, 254, 305–306, 311, 317–321, 323–324,
409, 412–414, 422–423, 431, 442, 485–491, 493–494, 496–511 328, 343, 356, 366
transport capacity 31, 50, 55, 171, 179, 243 wave ripple 46, 107, 260, 312, 331, 365, 370
transport distance 25, 412 wave-dominated 8–10, 171, 173, 177, 179, 183–186, 189–191, 193–
transport rate 27, 30, 50, 52, 57, 89, 91–95, 99, 135–136, 139, 141, 194, 196–197, 199, 200, 202, 207, 215, 225, 241, 249, 251–252,
145, 148–149, 179, 242 255–257, 260, 265, 267, 270–272, 281, 285, 293, 295, 300, 311,
transverse 22, 28, 32–35, 39–43, 45, 50–51, 53, 62–63, 67, 70, 72, 318, 344, 349, 363–364, 366, 374, 376, 381
95, 104, 139–140, 407–408, 425, 428, 437, 440, 446–447, 468, wave-influenced 237, 240, 242, 244, 250, 253, 255–256, 258–262,
472, 498 267, 270, 279–281
Triassic 43–44, 47, 60, 62, 65–67, 153 wavy laminae 32, 46
Tsondab Sandstone 63 wet grainfall 33
turbidite 2, 4–6, 8–9, 12–13, 203, 247, 250, 279, 370, 399–407, 412, wet interdune 38, 46, 51, 62–63, 66, 69, 71, 74–75
414–415, 417, 419–420, 423, 425, 427–428, 431, 434–436, 441– wetting front 32
442, 446–447, 458–459, 461–462, 466–468, 474–476, 480–484, Wheeler Gorge 8, 414, 447, 457–465, 476
493, 501, 508–509, 511 White Sands 19, 45, 50
turbidity current 356, 369, 373, 399–404, 407–409, 412, 415, 422, Wilmslow Sandstone 62, 65
425, 427, 429, 431, 435, 441, 447, 474,–476, 484, 493 wind gustiness 25
wind power 50, 52
wind regime 19, 22, 34, 40
U wind reversal 34
unidirectionality index 28 wind strength 27
unit bar 86– 87, 91, 93–102, 104–105, 107–110 wind velocity 22–23, 27, 37
United Arab Emirates 25 wind-ripple stratification 30, 36
universal facies 6, 13 Wingate Sandstone 62
upper shoreface 265, 300, 311–314, 317, 321, 323, 327, 329–332 winnowing 26, 243, 271
Utah 7–8, 36–38, 43–48, 58, 60, 62–63, 65, 68–70, 77, 186, 199, 238, wireline log 70, 73, 115–116, 197, 203, 209
243, 248, 262–263, 268, 271, 278–279, 283, 302–304, 321, 333, 373

Y
V Yellow Sands 62
valley 46, 87, 93, 119, 121, 134–139, 141, 143, 145, 147–148, 157–
160, 171–175, 177, 179–180, 182–185, 189–204, 206–212, 215–
221, 223–226, 237, 249, 253–254, 265, 267–270, 278–280, 282, Z
300, 341, 353, 372–374, 383, 389–390, 414, 493, 511 Zechstein Sea 62
valley fill 145, 147, 158–160, 171–172, 174, 180, 183, 191, 193–202, zibar 47
207, 211, 216, 219–220, 224, 253, 265, 267–268, 279–280
vegetation 22, 32–33, 43, 47, 50, 55, 63, 86, 91, 97–98, 103–104,
108, 110, 119, 121, 126, 134, 141, 147–149, 156–158, 270
velocity gradient 23
ventifact 26
Volga basin 269
528 INDEX

S-ar putea să vă placă și