You are on page 1of 152

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 exista 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 modelsindeed, 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 earths 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 Gresslys 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. 117.
2 ROGER G. WALKER

FaciesA 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 Potters
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 Potters
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 fansnevertheless, 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 Potters supply and other internal and external factors. Thus the various
conclusions remains true todaythe 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 facieshummocky 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 claysthe 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- tionsthe 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 parametersfor 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, Mialls (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 systemthey 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 environmentsmedium-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 successionfor 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 modeleddo 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 characteristicsthe 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 interpretationswe 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 usedreaders 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 entitiesfor 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 falla
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 modelsfor 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 preservedthe 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. Mialls 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 Mialls 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
todays 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 Allens 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 examplesis 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? Mialls (1977, 1985) fluvial depositional ele-
interpretationwas 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 modelan 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 importantthey 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. Barrierlagoon system rests disconformably on Pleistocene diamictite
(blue, conglomeratebreccia 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 transgressionit is washed into the lagoon
(dark red) during storms (blue arrow) and is also moved offshore to form small ridges (red arrowsouthern 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 Allens 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 barrierlagoon 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 Formationa model for deposition in shal- tion of Canada, Reprint Series 1, p. 1113.
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. 253279.
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. 1100.
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. 13731405.
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. 162.
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 261308.
elements) closely agree with the depositional elements defined MIDDLETON, G.V., 1973, Johannes Walthers law of the correlation of facies:
by modern processes. Geological Society of America, Bulletin, v. 84, p. 979988.
MIDDLETON, G.V., 1978, Facies, in Fairbridge, R.W., and Bourgeois, J., eds.,
ACKNOWLEDGMENTS Encyclopedia of Sedimentology: Stroudsburg, Pennsylvania, Dowden,
Hutchinson & Ross, p. 323325.
I thank Brian A. Zaitlin and Henry Posamentier for their MUTTI, E., AND RICCI LUCCHI, F., 1972, Le torbiditi dellAppennino
comments on the manuscript. I also thank the Natural Sciences settentrionale: introduzione allanalisi de facies: Societ Geologica
and Engineering Research Council of Canada for their support of Italiana, Memorie, v. 11, p. 161199. English translation by T.H.
my research. Nilsen, 1978, International Geology Review, v. 20, p. 125166.
POTTER, P.E., 1959, Facies models conference: Science, v. 129, p. 12721273.
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. 10531079.
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. 298323. ciation of Canada, p. 179194.
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. 237293. 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. 3147. 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. 102119. 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. 27182744.
Kinderscout Grit: a deltaic complex in the Namurian of northern VISSER, M.J., 1980, Neapspring 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- 543546,
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. 152. tary structures of turbidites. Yorkshire Geological Society, Proceed-
GRESSLY, A., 1838, Observations gologiques sur le Jura Soleurois: Neue ings, v. 35, p. 132.
Denksch. allg. schweiz., Ges. ges. Naturw., v. 2, p. 1112. 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. 737748.
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 GarringtonCaroline area: Bulletin of Canadian Petroleum
HARMS, J.C., AND FAHNESTOCK, R.K., 1965, Stratification, bed forms and flow Geology, v. 31, p. 213230.
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. 84115. with the Upper Cretaceous Wheeler Gorge conglomerates, Califor-
FACIES MODELS REVISITED 17

nia: a possible channellevee complex: Journal of Sedimentary


Petrology, v. 55, p. 279290.
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. 185272.
WILLIAMS, P.F., AND RUST, B.R., 1969, The sedimentology of a braided river:
Journal of Sedimentary Petrology, v. 39, p. 646679.
WILSON, J.E., 1975, Carbonate Facies in Geologic History: New York,
Springer-Verlag, 471 p.
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 Kuenens 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 unknownindeed, Kuenens 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. 397520.
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 paperGrowth 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. Normarks 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 (19701980). 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 limitationsthe 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 modelsthe 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, Rhne, 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 19321955, there have been 15 cable breaks,
study of the Rhne Fan, showing the existence of meandering averaging one every 1.5 years (Heezen, 1956). The flow of 1935
channels, channellevee 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 channellevee 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 DecemberJanuary and
The studies in the Bouma et al. (1985) compilation essentially AprilMay, 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 (18921903 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 todays 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
1937an 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 12 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 FACIESTHE 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 1960sthe 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. Lowes (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 successionHarms 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

outcropsconsequently 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 10100 cm in on the bed. Within the category of thin-bedded turbidites, the
thickness; Fig. 5) tend to be composed of Boumas 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 Boumas division A, rates of deposition from suspension during formation of the
and the sandstones contain only the BC 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 interpretationthere 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 sedimentwater 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 Boumas 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
abovegraded bedding, cross bedding, and imbricationall 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 FACIESOTHER 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
pebblesandmud 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 canyonsand 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 mapthis 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 magnitudethis attribute highlights the edges of the canyon floor as well as the edges of the tributary channels. C)
seismic reflection curvaturethis 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 = 200250m
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 shelfslope 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 fillthe 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 230280 m wide, with fills 1315 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 111), yellow (1335), and red deposition is only a few kilometers. Deposition may be strongly
(3638) 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 (1020 m deep, 200300 m wide) leveed chan-
tions in CC1 suggest flows from smaller tributaries along the nels can develop within canyon headsthis 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 1935 (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 modelsusing 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.
modernancient 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
slopethus 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. Orangechannel fills of unstratified bouldery
to pebbly conglomerate and very coarse-grained sandstone. Yellow and redunstratified coarse-grained sandstone and parallel
stratified medium-grained sandstone. Greeninterbedded 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. 6264). 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. 6669). 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 wallnote 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 sectionbut 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 channellevee 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 BC) 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 ExampleSimi 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. 7780). 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 channellevee 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
AF

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. AF) 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 thinthat
is, high on a channel margin, or on the back side of a levee. In Outcrop ExampleWheeler 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 channellevee 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: ~ 67 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 channellevee 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 parkit 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 channellevee


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 distaleither 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 sandstonemudstone 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 conglomeratesandstone 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 thicknesstwo 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 abovea 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 FailureNew 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, 110112). 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 degreesthis 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 channellevee 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 channellevee 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. 115120). 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 channellevee 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 smoothany 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 ExampleCounty 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 Rossthe 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 obstructionscharacteristic 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, slumpslide 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 bestKilbaha 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. 7173).
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 channellevee 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 scalethe 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. mentsnon-packaged turbidites, thickening-upward packages,
In the light of the seismic images presented here, particularly channel fills, and slumpslide 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 modelthe theme developed by Walker (this 145148). 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 DepositsMass-Transport Complexes or vertical section commonly is chaotic to contorted (Figs. 135,
150153), 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, 136140). 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. 3739) 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 changeprimarily 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 BD) 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 geometriesprocesses 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 differentfor 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 Rhne 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 datain 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 BE) 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 166168 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. 631648.
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. 466499.
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. 4060. 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. 153165. SEPM Foundation, 20th Annual Research Conference, p. 104115.
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. 426442.
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 channellevee com- Gulf Coast Section SEPM Foundation, 20th Annual Research Confer-
plexes, Early Eocene Regencia Canyon, Espirito Santo Basin, Brazil: ence, p. 127150.
Sedimentology, v. 44, p. 1746. 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. 7992.
15431560. 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 habitatschannel. 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 habitatslevee. 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. 9931009. Annual Research Conference, p. 402421.
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
470473. Water environments: London, Chapman & Hall, p. 196202.
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. 9498. of Deep Marine Clastic Systems: Geological Society of London,
DROZ, L., AND BELLAICHE, G., 1985, Rhne deep-sea fan: morphostructure Special Publication 94, p. 3149.
and growth pattern: American Association of Petroleum Geologists, KOLLA, V., AND COUMES, F., 1987, Morphology, internal structure, seis-
Bulletin, v. 69, p. 460479. 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. 650677.
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. 10891118. 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. 13731405.
Geological Society of America, Bulletin, v. 90, p. 11651176. 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. 14771488.
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. 91127.
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. 318329. 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. 248282. 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. 850873.
channel: Geological Society of America, Bulletin, v. 81, p. 13611384. 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. 157204.
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. 127133. Guidebook, p. 1729.
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. 279297.
Hydrodynamic Interpretation: SEPM , Special publication 12, p. MANLEY, L., AND FLOOD, R.D., 1998, Cyclic sediment deposition within
84115. 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. 912925.
Geogrfica de Colombia, Boletn, v. 51/52, p. 135143. 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. 11261149. 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. 533555
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-hydratecontrolled catastrophic sediment failures of the
Results, v. 155, p. 5378. Amazon fan: Geology, v. 26, p. 11071110.
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 Dor, 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. 578586.
ference, p. 133144. 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. 265293.
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. 387394.
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 channellevee systems of the Amazon Fan re-
Journal of Geology, v. 103, p. 285299. 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. 124137. Scientific Results, v. 155, p. 722.
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. 6593. 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) nellAppennino di Piacenza: Memorie POSAMENTIER, H.W., 2003a, A linked shelf edge delta and slope channel
dellAccademia 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. 141. 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. 75106. Bob F. Perkins Research Conference, p. 115134.
MUTTI, E., AND RICCI LUCCHI, F., 1972, Le torbiditi dellAppennino POSAMENTIER, H.W., 2003b, Depositional elements associated with a basin
settentrionale: introduzioni allanalisi de facies: Societ Geologica floor channellevee system: case study from the Gulf of Mexico:
Italiana Memorie, v. 11, p. 161199. Marine and Petroleum Geology, v. 20, p. 677690.
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 phyConcepts and Applications: SEPM, Concepts in Sedimentology
Petroleum Geology, v. 20, p. 523933. 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. 16871709.
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. 714738 (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. 127136.
Petroleum Geologists, Bulletin, v. 64, p. 10941101. 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 Iconceptual framework, in Wilgus, C.K., Hastings,
ciation of Petroleum Geologists, Bulletin, v. 54, p. 21702195. 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. 109124.
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. 367388.
Special Publication 68, p. 7388. 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 avulsionExample 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.
75174. 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 systemsUltra-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. 9399. 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. 806816.
Sedimentology, v. 30, p. 681694. 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., ChangesAn Integrated Approach, October 2023, 1985 (Abstract
Sedimentological Consequences of Convulsive Geological Events: and Poster).
Geological Society of America, Special Paper 229, p. 7792. 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. 701728.
Proceedings of the Ocean Drilling Program, Scientific Results, v. 155, READING, H.G., AND RICHARDS, M., 1994, Turbidite system in deep-water
p. 109146. 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. 792822.
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. 2345. Geologists, Bulletin, v. 50, p. 13891398.
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. 210. Springer-Verlag, p. 323347.
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 (6167N): 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. 910937. 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 Dor, A.G.,
Gulf Coast Section of SEPM Foundation, 20th Annual Research and Vining, B.A., eds., Petroleum Geology: North-West Europe and
Conference, p. 928939. Global Perspectives: 6th Petroleum Geology Conference, Proceed-
SURLYK, F., AND NOE-NYGAARD, N., 2003, A giant sand injection complex: ings, p. 199216.
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. 6981. 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. 17261752. 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. 10591086.
of sea level from coastal onlap, in Payton, C.E., ed., Seismic Stratig-
raphyApplications to Hydrocarbon Exploration: American Asso-
ciation of Petroleum Geologists, Memoir 26, p. 6381.
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. 527.
WALKER, R.G., 1965, The origin and significance of the internal sedimen-
tary structures of turbidites: Yorkshire Geological Society, Proceed-
ings, v. 35, p. 132.
WALKER, R.G., 1967, Turbidite sedimentary structures and their relation-
ship to proximal and distal depositional environments: Journal of
Sedimentary Petrology., v. 37, p. 2543.
WALKER, R.G., 1973, Mopping up the turbidite mess, in Ginsburg, R.N.,
ed., Evolving Concepts in Sedimentology: Baltimore, The Johns
Hopkins Press, p. 137.
WALKER, R.G., 1975a, Generalized facies models for resedimented con-
glomerates of turbidite association: Geological Society of America,
Bulletin, v. 86, p. 737748.
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. 932966.
WALKER, R.G., 1985, Mudstones and thin-bedded turbidites associated
with the Upper Cretaceous Wheeler Gorge conglomerates, Califor-
nia: a possible channellevee complex: Journal of Sedimentary
Petrology, v. 55, p. 279290.
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. 239263.
WEIMER, P., 1989, Sequence stratigraphy of the Mississippi Fan (Plio-
Pleistocene), Gulf of Mexico: Geo-Marine Letters, v. 9, p. 185272.
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, 133134, 179, 195, 201, 203, 207, 212,
A 249, 251, 253, 258259, 262, 265, 311312, 314, 318, 320, 336,
abrasion 23, 137 350, 363, 369, 371, 385, 389, 431
Abu Dhabi 40 Blackhawk Formation 303304, 321, 333
accommodation 22, 58, 60, 63, 171, 199, 211, 344, 350351 Bouma sequence 1213, 369, 399, 403404, 407
accumulation rate 52, 66, 71 boundary layer 2223
accumulation surface 32, 50, 5253, 55, 57, 6668, 74 bounding surface 13, 15, 1920, 22, 32, 34, 3840, 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, 9698, 102, 112, 138, 475
adhesion warts 32, 39 Brazil 62, 252, 256, 361, 427428, 430432
aerodynamic configuration 53, 55, 60 breaker bars 295, 299300, 311
Al Liwa sand sea 40 brinkline 31
algal growth 33 brittle failure 33
allocyclic 19, 175, 179, 189, 237, 281, 420 burrow 33, 4547, 65, 107108, 122123, 126, 129134, 193, 203,
allogenic 50, 60, 67, 73 219, 248249, 251, 259, 261, 265267, 280, 314, 329, 363364,
alluvial architecture 129, 141145, 147148, 150, 153156, 159 385
alluvial fan 24, 50, 71, 80, 85, 119, 135, 137, 139, 140141, 145, bypass 22, 50, 52, 55, 57, 5960, 66, 69, 71, 199200, 216217,
148, 156157 246, 267, 280, 344, 359, 365, 401402, 476, 483, 511
alluvial valley 87, 119, 138, 141, 145, 217, 249
anastomosing 96, 134135, 138, 139, 160
angle of climb 5253, 5556, 62, 6669, 73, 7577 C
angle of internal friction 33 Caliente Range 321, 331332, 337
angle of repose 27, 3132, 89, 91, 104, 122, 272 California 50, 249, 302, 304, 307, 319321, 323325, 328329, 331
anhydrite 65 332, 335, 337, 377, 405, 408412, 414, 429431, 433, 436, 446
animal activity 46 447, 454464
animal trackway (footprint) 33, 43, 46 canyon 8, 113, 279, 304, 361, 363, 373374, 399402, 415, 422
antidune 8586, 90, 95, 99, 104, 119 432, 454, 485, 493494, 504, 511512, 517
arid 19, 22, 24, 7071, 73, 107, 119, 124, 129130, 133, 139, 140, capillary fringe 45, 53, 55, 65, 68, 126
156157 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, 458459, 461, 474
Askja, Iceland 39, 44, 50, 55 Cedar Mesa Sandstone 3638, 4344, 4647, 58, 65, 6870, 77
Atchafalaya 241, 243, 246, 250, 272, 347, 352, 355, 381 cement 22, 55, 6265, 71, 73, 136, 220, 279, 314, 344, 350, 364, 409
Australia 5, 32, 44, 47, 65, 176177, 185, 221, 226, 249, 255, 340, cementation 63, 65
342, 360, 366, 369371, 390391, 403 channel 12, 57, 9, 12, 24, 63, 68, 71, 8587, 91, 93124, 126,127,
autocyclic 19, 34, 57, 175, 179, 243, 348, 359, 419420 129, 133148, 150151, 156160, 171, 176177, 179180, 184,
autogenic 73 186, 188196, 199200, 202204, 206208, 210211, 216217,
avalanche 27, 3032, 40, 104, 107 220, 225, 227, 237, 240241, 243244, 246247, 249250, 252
avalanche strata 3031 253, 257, 260, 262263, 265272, 274275, 279, 281, 285, 300,
avulsion 119, 121122, 134143, 147, 150151, 156160, 237, 243, 312313, 316317, 331, 353, 361, 367369, 372, 399404, 412,
249, 252253, 266267, 274, 285, 421, 441, 448, 463, 501 414451, 453454, 456459, 461, 463465, 468470, 472478,
482485, 486, 489490, 493, 495496, 501502, 504, 510511,
513514, 516
B channel belt 85, 87, 91, 97, 102, 108112, 115122, 126, 134139,
Bahrain 50 141144, 146, 151, 156, 158, 270, 438, 449, 476, 484
bar and trough system 316 channel fill 2, 6, 91, 96, 100101, 104, 107110, 112, 114, 117118,
barchan (barchanoid) 28, 32, 34, 4043, 53, 63 122, 133, 141, 159, 257, 265, 401, 425, 427432, 436, 440, 446,
barrier island 10, 200, 240, 244, 260, 271, 297298, 300, 303, 381, 389 449, 457458, 461, 463, 475476, 482484, 504
basalt 25, 6264 channel migration 86, 9697, 99, 103, 119, 137, 141, 176, 190, 253,
baseline of erosion 58 266, 434
bedform 1920, 22, 2730, 32, 3440, 43, 4553, 5557, 6264, 66, channel pattern 91, 93, 9699, 102103, 116, 118119, 134135,
68, 70, 7273, 76, 85, 8991, 102, 104105, 113, 174, 221, 243, 137138, 140, 143144, 147148, 156159, 186, 203, 253, 419,
270271, 295, 306307, 311, 313, 361, 363, 366, 371372, 385, 425, 427
403404 chenier 243, 268, 271272, 355
bedform behavior 34 climate 19, 22, 43, 4647, 58, 60, 63, 69, 71, 73, 80, 85, 107, 119,
bedform climb 22, 43, 50, 52, 5657, 62 121122, 124, 126, 130, 133135, 139, 141, 143, 147149, 151,
bedform migration 19, 22, 34, 39, 4849, 52, 56, 66, 68, 73 155160, 171, 178179, 199, 216, 225, 241, 254, 307
bedload 23, 8687, 89, 94, 9799, 104, 119, 137, 182183, 186, 191, climate change 22, 58, 60, 69, 71, 73, 122, 135, 141, 143, 148149,
243, 267, 295, 353, 361 151, 156158, 241
bedload sheet 8687, 89, 9899, 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. 521527.
522 INDEX

climb 3, 36, 40, 52, 53, 5556, 62, 6669, 7576 Devonian 5, 11, 44, 63, 112, 114115, 123, 129, 131, 157, 238, 253
clinoform 237, 239, 250, 262, 271276, 280, 362, 363, 367, 385 254, 274, 403, 406
coastal 1, 22, 27, 43, 4547, 50, 5758, 60, 62, 73, 119, 139, 147, 150, Dhahran 45, 50
157159, 171, 175177, 179182, 191193, 197, 215, 217218, diachronous 58, 199
221, 226227, 237, 240, 249, 254, 258, 293, 297298, 300, 303, diagenetic cement 55
305306, 318320, 323324, 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, 202203, 237, 241, 243244, 246, 249
cohesive 22, 32, 44, 97, 425 250, 252253, 257, 260, 262263, 265272, 274275, 279, 417,
cold-climate desert 47 422, 427, 431, 463, 468, 473
Colorado 11, 40, 60, 62, 201202, 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 2731, 35, 43, 51, 6263
compound bar 86, 91, 94, 98, 100102, 104105, 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, 275277, drift direction (resultant, RDD) 28, 34
304305, 319, 363, 367, 379, 381382, 386, 407, 427, 431, 452 dry eolian system 19, 5253, 55-56, 6062, 68, 74
crawling trace 33, 129 dry interdune 31, 43, 45, 66, 68, 78
creep 23, 2627, 37 Duero Basin 46
Cretaceous 34, 6, 912, 36, 6263, 153, 157, 174, 177, 180, 184, dune 13, 1920, 2223, 25, 2736, 3848, 5057, 60, 6263, 6576,
186, 191192, 196197, 199200, 202, 204, 206208, 237, 243, 78, 8591, 9495, 98100, 104, 107, 109, 112, 115116, 119, 121
248, 253, 261263, 266268, 270271, 273274, 276281, 283, 122, 139, 187, 191, 227, 268, 295, 312, 316317, 319, 329, 361,
285, 302303, 333, 341, 374376, 405, 407, 410411, 414, 427, 365366, 370, 372
430, 446447, 454, 501 dune element 40, 43, 4546, 54, 6263, 6667, 69
crevasse splay 91, 115, 119124, 126127, 134135, 137, 139, 270, dune flank 27
400, 422, 438, 463, 468474, 504 Dunvegan Formation 180, 191, 248, 261, 266, 268270, 274, 276
crinkly laminae 47 282
critical climbing 52 dynamic facies model 19, 22, 67, 69, 73, 77
cross bedding (cross-bedding) 2, 68, 1011, 19, 32, 34, 3839, 42,
190, 249, 258, 261, 265, 298, 307, 311313, 317, 319, 322323,
331, 363364, 366, 371, 373, 375, 380, 389, 407 E
cross strata 32, 3335, 3841, 4850, 52, 55, 63, 86, 8990, 94, 98 Ebro delta 252
100, 104, 107, 109, 112, 121122, 127, 268 ejecta 23
cross stratification 7, 13, 15, 32, 78, 88, 104, 195, 249, 258, 260, 281,
England 3, 10, 13, 4344, 47, 62, 6566, 191, 215, 364, 413
381 Entrada Sandstone 48, 62, 65
crust 33, 46, 129, 134, 148, 149, 153, 157, 353 entrainment 19, 2223, 25, 50, 53, 119, 351
curl 33, 45, 46 eolian 6, 8, 13, 1920, 22, 23, 2540, 4348, 50, 5258, 6063, 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, 5557, 6162, 6869, 73, 75
D eolian system, wetdry 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, 259260, 267, 270 180, 237
deflation 19, 22, 26, 28, 43, 4748, 52, 55, 5761, 6566, 6970 erg 22, 24, 4546, 50, 53, 5758, 60, 62 63, 65, 6667, 6971, 75
delta 23, 6, 910, 13, 16, 50, 85, 107, 115, 119, 121122, 134135, 77, 80, 88, 143145, 176, 180, 239, 257, 272, 297
138139, 143, 153, 157158, 173174, 176, 178180, 182, 184 erg construction 22, 50
185, 189191, 194, 196, 199204, 206208, 210, 215220, 225, extra-erg 46
237268, 270282, 284285, 293, 320, 324, 339, 342, 344345, estuarine lithosome 189, 190
347, 349350, 352353, 355, 360, 363, 367, 371373, 375, 381 estuary 5, 171, 173, 176177, 180, 182191, 193, 195, 198201, 206
383, 390, 401, 425, 427429, 431, 436, 463, 471, 493 208, 212, 215221, 225226, 237, 256, 265, 268, 282, 285, 344,
delta front 184, 246247, 249251, 253, 256, 257, 259, 261262, 363, 372, 401402, 445
266268, 270272, 279 Etendeka igneous province 62
delta, lacustrine 115, 119, 121122, 139, 249, 276 Etjo Formation 36, 6263
delta plain 157158, 184, 237, 244, 249250, 253257, 265, 270, eustasy 147, 158, 254, 281, 297, 339
272, 355, 381, 431, 493 evaporite 33, 47, 65, 124, 129, 156157
density 2223, 37, 89, 129, 153, 179, 203, 241, 243, 246, 258, 347, evaporite precipitation 33, 47
369, 370, 401403, 408, 415416, 421, 425, 427, 493, 504
density current 369, 401
depositional elements 2, 911, 13, 1516, 24, 183, 399401, 403, F
414415, 420, 422, 428, 431, 447, 470, 484485, 501, 504 facies architecture 19, 69, 172, 237238, 240, 255, 258, 264, 268,
depositional model 64, 6668, 85, 86, 99100, 111, 119, 121124, 270, 281, 388
135, 171, 293, 355, 414 facies association(s) 1, 57, 13, 15, 54, 71, 298299
desert 13, 19, 2223, 33, 3940, 4547, 5051, 58, 63, 66, 129 facies definition 2
desert rose 33 facies model(s) 13, 610, 1213, 1516, 19, 22, 39, 6667, 6970,
desiccation cracks 11, 33, 4546, 57, 107, 119, 122, 126, 133, 156 72, 73, 75, 77, 85, 171172, 174180, 182183, 189, 191, 212,
INDEX 523

215218, 220, 223, 225, 237, 282, 293298, 300, 302, 304305, Great Sand Dunes 40
311, 322, 324, 339, 341, 345, 347, 363, 369, 379380, 390, 399, ground-penetrating radar 22, 85, 179, 270
430, 504 growth fault 243, 256259, 263, 274, 279
facies modeling 13, 13, 16, 6667, 177, 430 Guerrero Negro 46, 58
facies succession(s) 12, 67, 71, 177, 179, 183, 215, 237, 240, 255, Gulf of Mexico 9, 147, 200, 237, 239240, 243, 253, 272, 275, 305,
257258, 260261, 264, 267, 272, 277281, 304, 504 339, 341, 348349, 352, 356, 369, 372, 378, 383, 390, 423, 426
fair weather (fair-weather) 9, 15, 250251, 294, 296, 300, 311314, 427, 429, 431, 437, 439, 441443, 446447, 449, 453454, 469
317320, 323324, 326, 328, 336, 356, 360, 369, 377 470, 475476, 478, 485486, 489, 492, 494, 498, 501502
fenestral porosity 33 gutter casts 317, 323, 367, 374
Ferron Sandstone 262263, 268, 271, 283 gypsum 33, 46, 50, 65, 129
flake 33, 46
flash flood 46
flood basalt 62, 64 H
flood basin 115, 117, 119121122, 126, 134, 136137 halite 33
flood deposit 107, 137, 255256, 259, 377 hard-pan 46
flood plain 8587, 96, 110, 113, 115, 117123, 126, 129130, 134 Helsby Sandstone 4344, 62, 6668
139, 141145, 147, 150, 153, 156157, 159160, 179, 186, 192, high-energy coast 293, 304, 311314, 319320, 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, 370371, 373374
fluid mud 243, 246, 253, 311, 345, 355 hyperpycnal 241, 246247, 249, 253, 258, 282, 344, 347348, 350,
fluvial 2, 6, 911, 15, 19, 27, 43, 46, 50, 6263, 6566, 6871, 8588, 353, 356, 358, 370, 374, 402, 415, 427, 493
111, 113, 129, 131135, 141144, 146150, 153, 156160, 171 hypopycnal 241, 243, 249,250, 258, 344345, 348, 352353, 355,
172, 174176, 178180, 182183, 186, 189192, 194, 195202, 358
204205, 207, 209212, 215218, 225226, 237, 240, 242, 249,
253, 255256, 258, 260262, 265, 270, 272, 277, 279282, 331,
339, 344, 347348, 350, 353354, 359, 415, 425, 427, 434435, I
444 Iceland 29, 36, 39, 44, 47, 50, 55
fluvial inundation 46 ichnofacies 129130, 193, 195, 219221, 251, 259, 265, 280281,
fog 65 300, 302, 305, 369, 373, 375, 377, 381382, 385386
forced regression 257, 260, 279, 293, 299301, 318, 320322, 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, 159160, 171175, 177, 179180, 184, 191193,
foreset 10, 13, 30, 39, 41, 43, 47, 4950, 238239, 249250, 272274, 195201, 206, 208, 215217, 219220, 223, 253254, 279280,
320, 329331 341, 372, 374, 389, 493, 511
foreset azimuth 39, 43, 49 inclined strata 8687, 99105, 107108, 110111, 113114
fossil 45, 15, 47, 123, 129134, 157, 182, 190, 203, 207, 209, 212, Indian Ocean 360
215, 218219, 249, 300, 304, 359, 379 infiltration capacity 46
friction 2223, 33, 112, 183, 186, 241, 243, 246, 249, 253, 258, 267, interdistributary bay 3, 243, 249, 253, 268, 270
270, 350, 356, 360 interdune 1920, 22, 27, 3133, 38, 40, 43, 4548, 5157, 60, 6263,
frontal splay 250, 270271, 399400, 415417, 419422, 447, 473 6571, 7376, 78
478, 484485, 493, 504, 508, 511 interdune flat 38, 43, 45, 5253, 55, 57, 62, 6869
Frontier Formation 261, 273274, 277, 279, 284 interdune hollow 38, 51, 56, 6668
interdune migration surface 38, 48, 56
interglacial 43, 156158, 201, 346
G intra-erg 46, 63
Gallup Sandstone 265 inverse grading 27, 30, 37, 247, 403, 407
gas 1, 3, 19, 7071, 79, 85, 123, 129, 133134, 157, 159, 202, 210, isochronous 58
221, 237238, 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, 156158, 201, 246, 293, 302, 346,
366 J
Gondwana 62 Jackson Group 320, 328, 330
graded bedding 258, 399, 404, 407408 Jafurah area 46
grain packing 32 Jurassic 48, 60, 62, 65, 133, 153, 354
grainfall 3033, 36, 38, 4041, 78
grainfall strata 3032, 40
grainflow 27, 3132, 37, 4041, 4344, 78 K
grainflow, amalgamated 31 Kelso Dune Field 50
grainflow strata 3132, 44, 78 Kuiseb River 27
Gran Desertio 40 Kuwait 50
524 INDEX

Mekong 243, 251, 253, 261, 270


L Mexico 9, 19, 40, 4546, 50, 62, 147, 157, 200, 237, 239240, 243,
lacustrine 27, 50, 107, 115, 119, 121122, 124, 127, 129130, 139, 253, 265, 272, 274v275, 298, 305, 309, 339, 341, 348349, 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, 453454, 469, 470, 475476,
lag 2, 7, 915, 26, 28, 30, 43, 47, 50, 58, 63, 91, 148149, 172173, 478, 485486, 489, 492, 494, 498, 501502
180, 182, 193, 201204, 217, 221, 237, 239240, 249, 256, 260, migration 1920, 22, 2728, 30, 32, 3435, 3740, 43, 4849, 52,
264, 267, 270, 275, 280, 284, 297, 314, 320, 322, 349, 353, 365, 5556, 58, 6263, 66, 68, 70, 7375, 85, 86, 8991, 9697, 99,
368369, 402, 408, 430, 432, 493 101105, 109112, 115, 119, 121122, 124, 137, 141, 143, 148,
Lagniappe delta 275 159, 176, 179180, 184, 190191, 210, 243, 253, 266, 268, 272,
lagoon 7, 910, 1315, 173, 180, 182, 215, 217, 237, 239240, 242, 279, 297, 300, 306307, 313, 362363, 372, 379380, 383, 385,
249, 256, 260, 264, 267, 270, 493 387, 390, 424, 428, 434, 435, 439, 441444, 446, 451452, 475,
lagoonal 10, 1315, 237, 240, 264, 267 483
lake 24, 3233, 4547, 50, 5758, 60, 62, 71, 80, 102, 113, 119, 121 migration direction 28, 34, 3940, 48, 49
124, 127, 129130, 134135, 137139, 143, 153, 156157, 180, migration speed 34
184185, 191, 197, 201202, 206, 238239, 246, 249, 253, 414 Milankovitch 58, 62, 157, 158
Lake Lucero 50 mineralogy 22, 157, 379, 388
laminar flow 25 Mississippi 89, 9697, 116, 121, 157159, 174, 176, 180, 191, 197,
lateral accretion 434435, 441, 443446, 449, 483 201, 237238, 240, 242243, 246, 249250, 253254, 258, 260,
lee 20, 2728, 3135, 3739, 48, 5253, 91, 104, 216, 305, 311, 315, 264, 270, 272, 285, 344, 347, 349, 352, 355, 371, 373, 381382,
363 390, 401, 422427, 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 12, 6, 9, 12, 91, 115, 119124, 126, 134, 137, 139, 179, 189, moisture 22, 43, 45, 5253, 65, 126, 130
199, 243, 270, 399401, 404, 406, 412, 414422, 425, 427, 429, mojave 22, 43, 45, 5253, 65, 126, 130
431432, 437443, 446447, 449451, 453, 456, 458459, 461, morphology 19, 27, 3839, 42, 45, 50, 55, 62, 6667, 70, 73, 171,
463464, 468470, 472478, 485, 490, 493, 494, 496, 501502, 174176, 178, 182184, 215, 227, 237238, 240242, 253, 255
504, 510511, 513514 256, 316, 323, 339, 390, 417, 419, 422, 428, 471472, 477, 501,
leveed channel 1, 400, 401, 415417, 419422, 427, 429, 431, 437 504
438, 441443, 447, 450, 473478, 485, 493, 496, 504, 511 mud curl 4546
liquefaction 3233, 43 mud drape 46, 50, 107, 195, 198, 203, 205, 262, 267, 361, 373, 375
lobe 1, 2, 91, 9697, 104, 119, 139141, 158, 237, 240242, 244, 246, mud flake 33, 46
250, 255256, 260, 262, 264, 267268, 270271, 274275, 277 muddy substrate 46
281, 340, 343, 381, 390, 399401, 415418, 420, 422, 473, 475
476, 484, 496, 501
loess 23 N
low accommodation 175, 191, 199, 202, 209 Namib Desert 45, 47, 5051, 63
low-energy coast 317319, 324, 333 Namib Sand Sea 27, 32, 35, 4546, 50, 52
Lower Cutler Beds 36, 45, 63, 68 Namibia 27, 29, 32, 3637, 4546, 55, 58, 6263
lower shoreface 300301, 318, 320323, 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, 316317, 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, 159160, 199, 297, 300, 323 North Sea 19, 58, 62, 70, 7881, 358, 375
mass-transport complex 401, 412413, 485491, 493, 498499, 504, numerical modeling 220221
508
Mauritania 43
meandering 7, 9, 12, 15, 91, 9698, 101102, 108, 112, 116117, O
135, 138, 140, 142, 156, 175, 186, 188, 190191, 196, 199202, oblique migration 35, 63
209210, 212, 268, 401, 427, 434435, 444, 501 Oman 25
meandering channel 97, 101, 188, 401, 434435, 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, 306307, 311 312, 315317, 319, 323, 327, Oregon coast 45, 312, 320
370 Ormskirk 62, 68
INDEX 525

overbank 71, 94, 107, 115, 119, 121, 124, 136138, 140, 143, 156, ravinement 189191, 194, 199201, 206, 209, 218, 224, 251, 279
158, 160, 194, 199200, 210211, 415, 417, 422, 438, 441442, 280, 297, 300, 349, 351, 372, 376, 381382, 386, 388
450, 452453, 463, 468, 473474, 476, 482484, 501, 504 reactivation surface 3435, 38, 48, 65, 195, 261262, 375
recognition criteria 197, 217, 361
reg 26
P regional climate 46, 58, 157
Padre Island 41, 4546, 305, 308, 320, 326 regression 15, 60, 113, 149, 159, 180181, 192193, 197, 215, 257,
Page Sandstone 60 260, 277, 279, 293, 296, 297, 299301, 318, 320322, 324, 337,
paleoenvironment 19, 57 349, 354, 431, 436, 493
paleosol 33, 46, 68, 113, 126, 128130, 156160, 193194, 196, 201 remote sensing 85
202, 209211, 267 reptation 23, 2627
parabolic 33, 40 reservoir prediction 70, 73
Paradox Basin 63 resultant drift direction (see drift direction, resultant)
Paraba do Sul 252, 259, 271 resultant drift potential (see drift potential, resultant)
Paran Basin 62 reversing dune 40
parasequence 196197, 202, 268, 279, 294, 297, 300, 321, 383, 386, rhizoliths 33, 47, 5758, 60, 69, 133134
388 Rhone (Rhne) (River, delta, fan) 8, 259260, 274, 401, 465, 501
particle size 22, 25 rip channel 312, 316317
pebble lag 7, 43, 193, 201202 rip current(s) 178, 293, 295, 311, 313314, 317, 320321, 357, 366
Pennsylvanian 3, 36, 63, 68, 180, 253254, 272, 274 ripple 34, 6, 9, 1113, 2733, 3641, 43, 4547, 52, 56, 78, 8587,
perched water table 46 8990, 9495, 98100, 104, 107, 112, 119, 122123, 126127,
permafrost 63 191, 203, 207, 210, 248249, 258, 260262, 298, 301, 306307,
permeability 46, 85, 107108, 110, 144, 201205, 207, 210, 211 311320, 323, 327, 331, 363365, 369370, 373375, 399, 403
212, 404 406, 427, 442, 446447, 458459, 467468, 473, 481
Permian 8, 19, 3638, 4347, 58, 60, 6263, 65, 6871, 7781, 180 ripple indices 27
pinstripe lamination 31, 36 ripple strata 3031, 3638, 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, 133134, 156 river mouth 176177, 189, 199, 215216, 218, 226, 240, 249, 253,
playa 45, 47, 71, 139 255, 257258, 267, 282, 304, 344345, 347, 350, 356, 402, 432,
Pleistocene 4, 14, 43, 113, 157158, 198, 223, 238, 320, 323324, 493, 511
328, 329, 335, 349, 375, 379, 402, 429, 437, 449 river-influenced 9, 240
Pleistocene terrace deposits 328329 root structure 33, 45, 47, 63, 330
plinth 27, 32, 36 Ross Sandstone 434, 445, 474, 476, 479480, 483484
Po delta (Po River) 243, 247, 252 Rotliegend 19, 58, 7071, 73, 78, 8081
point bar 1, 12, 91, 94, 96, 98102, 104, 109, 118, 186, 212, 434
435, 445
polygon 33, 57, 60 S
polygonal fracture 57, 60 sabkha 24, 47, 50, 55, 60, 62, 7172, 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, 248249, 251, 253, 255, 257 salt-growth 47
258, 260, 262264, 267268, 270272, 274, 280, 347 salt precipitation 47
progradational shoreline 297, 299 saltation 23, 27, 30, 3738
prograding strand plain 298 saltation cloud 23
Proterozoic 60, 65, 186 saltation path length 27, 37
pseudomorph 33, 46, 65 San Clemente 429431, 433436, 501
pumice 25 San Miguel Formation 280, 285
sand sea 20, 22, 27, 32, 35, 40, 4546, 50, 52, 40
sand transport direction 2728, 33
Q sand-accumulating bedform 28
Quaternary 22, 43, 58, 73, 134, 156158, 180, 199, 210, 239, 272, sand-transporting bedform 28, 34
339, 341, 343, 348, 354, 359, 363364, 371374, 375, 381, 385, sandflow 31, 34
388, 389, 390, 402 sandsheet 29
Queensland 65, 371 satellite 22, 24, 85, 221, 340342, 352, 369, 391
Saudi Arabia 4546, 50
scarp recession grainflow 31
R scour pit 39, 48
raindrop imprint 33, 122 sea level 2, 7, 910, 13, 22, 50, 58, 73, 122, 124, 141, 148, 150, 152
rainfall 3033, 36, 38, 4041, 46, 78, 130, 149 153, 156160, 171, 173175, 179182, 184185, 189192, 197
526 INDEX

202, 215218, 220, 224225, 227, 238239, 246, 254256, 279, slab slide 33, 37, 44
281282, 285, 293294, 296297, 299303, 306, 324, 337, 339, slipface 27, 28, 3132, 40, 6263, 127
346349, 354, 359, 390, 399, 401402, 425, 431, 436, 490, 493, slope channel 399400, 422, 425, 427429, 431, 437, 441, 444, 451,
501, 511 490, 504
seabed imagery 221 slump degradation grainflow 31
seasonal 27, 34, 46, 68, 119, 121, 129130, 133, 156, 241, 249, 360, slumping 32, 412, 425426, 441, 466, 474
363 soft-sediment deformation 32, 4344, 258, 260, 274, 458459, 461
secondary airflow 27, 35 soil 119, 126, 128130, 133, 138, 156, 160, 193, 353
sediment availability 19, 50, 6870, 73, 7677, 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, 7881, 358, 375
sediment budget, neutral 57 Spain 4546, 252, 319320, 324, 327329, 335
sediment compaction 58 spatial complexity 73
sediment flux 22, 57, 67, 149, 179180, 182183, 218, 220, 504 sphericity 25
sediment supply 2, 910, 13, 19, 22, 27, 30, 50, 55, 6465, 69, 85, staging area 399400, 415, 417, 422, 425, 485, 491, 493, 504
91, 122124, 135, 137, 138, 141, 144145, 148149, 151, 156 static facies model 67, 174
159, 171, 179182, 191192, 197, 199200, 216, 225, 250251, stepped transgression 297, 300, 321
255, 258, 279, 281, 297, 303, 341, 344, 347, 349350, 355356, Stokes surface 19, 57, 59
374, 380, 390, 422, 501 storms 14, 86, 90, 139, 241, 247, 249, 253, 258, 293, 295, 313314,
sediment transport 23, 50, 52, 8586, 9195, 98100, 107, 112, 119, 317319, 339, 350, 356357, 359360, 366367, 369370
135139, 141, 145, 147149, 158, 176, 179, 186, 217, 221, 242, stoss 27, 29, 3637, 52, 91, 305
258, 295, 320, 343, 350352, 360, 362, 366, 369, 370, 511 strataset 8587, 9091, 102, 108, 110, 115, 121123, 137
sediment transport rate 50, 52, 9193, 95, 99, 135,136, 139, 141, stratigraphic modeling 22
145, 148149, 179, 242 stratigraphic organization 175, 198, 200, 281
sediment waves 400, 422, 437, 446, 450451, 453454, 463, 468, subsidence 11, 46, 5758, 61, 65, 67, 119, 135136, 138139, 141
472, 475, 504 142, 147151, 153, 156, 159, 237, 257, 274, 279, 282, 297, 302,
sedimentary basin 137, 141, 147149, 180, 202, 211, 219, 220 339, 342, 381
seif 28, 144, 146 superimposed bedform 20, 30, 35, 39, 40, 48, 102, 104105, 366,
seismic 12, 5, 70, 85, 115116, 118, 141, 144145, 159, 171, 175, 385
179, 192, 197, 203, 206, 216, 218, 220221, 223, 238, 239, 256 superimposition 27, 35, 38, 40, 43, 48, 62
257, 267, 270, 272, 274275, 305, 362364, 367368, 372, 375 supersurface 22, 48, 55, 5760, 62, 66, 67, 6971
376, 383385, 399, 401, 413, 415, 417, 422426, 428429, 431, supratidal flat 65
434, 437444, 446447, 458, 468470, 473477, 484486, 489, surface creep 23, 26
490, 496498, 500501, 503508, 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, 5758, 60, 6669, 71, 73, swaly cross-stratification 301, 311
77, 9192, 100101, 103104, 107, 115116, 122, 124, 126127, systems tracts 193, 199, 216218, 238, 281282, 339, 348349, 359,
129130, 134, 137, 141145, 156, 157, 158160, 171, 175, 179, 386, 388, 390, 425
189, 191200, 202, 209, 211, 215216, 218219, 221, 237239,
241, 253, 255256, 260, 266, 268, 274, 279, 281283, 294295,
297, 299300, 318, 320322, 325, 328339, 341, 348, 351, 369, T
376, 383, 385389, 399401, 403404, 407, 414, 420, 431, 457, tectonism 58, 8586, 119, 122, 124, 135138, 141, 143144, 151,
459463, 475, 483, 485, 493, 510 156, 158160, 293
sequence boundary 57, 189, 191194, 197200, 202, 209, 215216, teepee 33, 47
218, 348, 388389 Tnr 47
sequence stratigraphy 159, 171, 175, 239, 274, 281282, 299, 341, Tertiary 63, 132, 134, 148, 153, 157, 359, 383, 431, 458, 501
385386, 401, 485 Texas 41, 4546, 180, 200201, 238, 249, 259, 281282, 285, 308, 310,
serir 26 312, 320, 324, 326, 328, 330, 336, 349, 353, 369,371, 373, 378
setdown 295, 299, 321 tide 3, 810, 171172, 176179, 181184, 186, 189191, 196, 197,
setup 295, 299, 321, 350, 357, 366367 215, 217218, 225, 237241, 243, 249256, 261262, 266267,
shadow zone 27, 37 270, 273, 277, 279280, 282, 285, 293, 307, 344, 349350, 359
shallow marine (see marine, shallow) 364, 366, 372, 375, 380
shear stress 23, 25, 8889, 9195, 112, 119, 137, 139, 243, 350, 360, tide-dominated 8, 171, 177179, 183184, 186, 189191, 196, 215,
368 217218, 237, 240241, 252254, 256, 267, 270, 277, 282, 344,
Shikaoda Formation 65 362363, 372
shoreface 180, 184, 189190, 196, 199200, 202, 207208, 210, 215, tide-influenced 239, 241, 251, 254, 256, 261262, 266, 270, 273,
217218, 221, 224, 237, 249251, 258260, 265, 267, 271272, 279280
278279, 281282, 293305, 307314, 317324, 327, 329333, top-truncated delta 251, 279, 281
339341, 349350, 355, 357, 366367, 370374, 380381, 385, topset 238239, 249, 251, 260, 272, 279, 281
390 trackway 33, 43, 4647, 134
shoreface profile 299300, 304, 308309, 323, 327, 339 transgression (transgressive) 10, 1415, 58, 60, 62, 149, 156, 159
shoreface-attached ridges 305, 310, 370372 160, 171172, 176, 179182, 184, 189192, 194202, 207, 210,
Silurian 63, 157, 179 215218, 224, 237238, 240, 251, 255256, 260, 265, 277278,
Skeleton Coast 27, 29, 32, 46, 55 280282, 284285, 293294, 297, 299300, 303305, 308, 321,
INDEX 527

323, 332, 339, 343344, 348349, 351, 354, 359, 362364, 367,
371, 374, 379, 380381, 383, 386388, 390, 425, 485, 493 W
transition point 417418, 420, 473, 475476, 485, 493, 508 Waitemata Group 415, 461, 465466, 468
translatent rippleform stratification 30 Walthers Law 237, 293
transport 2, 19, 2223, 2528, 3031, 3335, 38, 40, 45, 47, 5053, water flow 86, 91, 9394, 99, 107, 119, 137, 312
55, 57, 63, 67, 8586, 89, 9195, 97100, 107, 112, 119, 122, water table 19, 22, 33, 4547, 50, 52, 53, 5563, 6570, 7475, 77,
129, 133139, 141, 145, 147149, 15859, 171, 176, 179180, 85, 126, 130, 133, 138, 160, 193
183, 186, 196, 200, 217, 221, 225, 242243, 253, 258, 295, 298 water-table elevation 46
299, 303, 306, 311, 314315, 320, 343345, 348, 350353, 356, wave base 15, 294296, 318, 328, 369, 399, 493, 511
359363, 365366, 369371, 373374, 377, 390, 399, 401, 407 wave energy 183184, 250, 254, 305306, 311, 317321, 323324,
409, 412414, 422423, 431, 442, 485491, 493494, 496511 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 810, 171, 173, 177, 179, 183186, 189191, 193
transport rate 27, 30, 50, 52, 57, 89, 9195, 99, 135136, 139, 141, 194, 196197, 199, 200, 202, 207, 215, 225, 241, 249, 251252,
145, 148149, 179, 242 255257, 260, 265, 267, 270272, 281, 285, 293, 295, 300, 311,
transverse 22, 28, 3235, 3943, 45, 5051, 53, 6263, 67, 70, 72, 318, 344, 349, 363364, 366, 374, 376, 381
95, 104, 139140, 407408, 425, 428, 437, 440, 446447, 468, wave-influenced 237, 240, 242, 244, 250, 253, 255256, 258262,
472, 498 267, 270, 279281
Triassic 4344, 47, 60, 62, 6567, 153 wavy laminae 32, 46
Tsondab Sandstone 63 wet grainfall 33
turbidite 2, 46, 89, 1213, 203, 247, 250, 279, 370, 399407, 412, wet interdune 38, 46, 51, 6263, 66, 69, 71, 7475
414415, 417, 419420, 423, 425, 427428, 431, 434436, 441 wetting front 32
442, 446447, 458459, 461462, 466468, 474476, 480484, Wheeler Gorge 8, 414, 447, 457465, 476
493, 501, 508509, 511 White Sands 19, 45, 50
turbidity current 356, 369, 373, 399404, 407409, 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, 93102, 104105, 107110 wind velocity 2223, 27, 37
United Arab Emirates 25 wind-ripple stratification 30, 36
universal facies 6, 13 Wingate Sandstone 62
upper shoreface 265, 300, 311314, 317, 321, 323, 327, 329332 winnowing 26, 243, 271
Utah 78, 3638, 4348, 58, 60, 6263, 65, 6870, 77, 186, 199, 238, wireline log 70, 73, 115116, 197, 203, 209
243, 248, 262263, 268, 271, 278279, 283, 302304, 321, 333, 373

Y
V Yellow Sands 62
valley 46, 87, 93, 119, 121, 134139, 141, 143, 145, 147148, 157
160, 171175, 177, 179180, 182185, 189204, 206212, 215
221, 223226, 237, 249, 253254, 265, 267270, 278280, 282, Z
300, 341, 353, 372374, 383, 389390, 414, 493, 511 Zechstein Sea 62
valley fill 145, 147, 158160, 171172, 174, 180, 183, 191, 193202, zibar 47
207, 211, 216, 219220, 224, 253, 265, 267268, 279280
vegetation 22, 3233, 43, 47, 50, 55, 63, 86, 91, 9798, 103104,
108, 110, 119, 121, 126, 134, 141, 147149, 156158, 270
velocity gradient 23
ventifact 26
Volga basin 269
528 INDEX

You might also like