You are on page 1of 18

The American Association oi PelroleLim Cieologisls Bulletin

V, 73, No. 2 (February 1989), P. 125-U2, II Fig.s.

Genetic Stratigraphic Sequences in Basin Analysis I:


Architecture and Genesis of Flooding-Surface Bounded
Depositional Units^
WILLIAM E. GALLOWAY'

ABSTRACT
Marine basin margins are characterized by repetitive
episodes of progradation punctuated by periods of transgression and flooding of the depositional platform. The
resultant stratigraphic units consist of genetically related
(1) depositional systems and their component facies
sequences; (2) bypass, nondepositional, and erosional
surfaces; and (3) in thick sequences affected by gravity
tectonics and crustal response to loading, syndepositional structural discontinuities. Units are bounded by
hiatal surfaces preserved as submarine unconformities or
condensed sedimentary veneers and that record maximum marine flooding of the basin margin. The repetitive
stratigraphic architecture is the product of the ongoing
interplay among sediment supply, basin subsidence (and
uplift), and eustatic sea level change. Each of these three
variables may dominate depositional evolution; furthermore, stratigraphic architecture is very similar regardless
of the dominant control.
A genetic stratigraphic sequence is the sedimentary
product of a depositional episode. The sequence incorporates and reconciles depositional systems, bedding geometries, and bounding surfaces within the framework of
cycles of basin-margin offlap and flooding. Each
sequence consists of the progradational, aggradational,
and retrogradational or transgressive facies deposited
during a period of regional paleogeographic stability.
The defining genetic stratigraphic sequence boundary is a
sedimentary veneer or surface that records the depositional hiatus that occurs over much of the transgressed
shelf and adjacent slope during maximum marine flooding. The genetic sequence paradigm emphasizes preserving the stratigraphic integrity of three-dimensional
depositional systems and does not rely on widespread
development of subaerial erosion surfaces caused by
eustatic falls of sea level to define sequence boundaries.
The physical stratigraphic record of transgression and
Copyright 1989. The American Association of Petroleum Geologists. All
rights reserved.
^Manuscript received, June 10,1987; accepted, Augusts, 1988.
^Department of Geological Sciences, University of Texas at Austin, Austin,
Texas.
This paper is based In part upon research supported by National Science
Foundation grant EAR-841613B. The ideas evolved from a 1984 AAPG Distinguished Lecture series. Figures were drafted by Jeff Horowitz. Betty Kurtz
typed the bibliography. I thank William Bazeley, Franl< Brown, Nicholas
Christie-Blick, RobertDott.Jr, William Dickinson, Martin Lagoe, AndrewMiall,
and Don Swift, as well as reviewers L. L. Sloss and J. F. Sarg for their comments, critique, and support. I dedicate this paper to Dave Frazier, who taught
me to worry more about genetic stratigraphy than he probably guessed.

floodingdistinctive thin but widespread facies


sequences, prominent erosional surfaces, and superjacent marine condensed intervals or sedimentary
veneersprovides readily recognized, regionally correlative, easily and accurately datable, and robust sequence
boundaries that commonly define times of major basinmargin paleogeographic reorganization in terrigenous
clastic basins.

INTRODUCTION
The successful interpretation and three-dimensional
delineation of depositional systems and basin-scale facies
tracts require the recognition and correlation of genetic
stratigraphic packages. Our final interpretations of
paleogeography, sediment dispersal patterns, and basin
history can be no better than the stratigraphic foundation
upon which they are based. Thus, establishing genetic
stratigraphic units for further mapping and analysis is a
fundamental but commonly difficult starting point in
basin analysis.
One potential approach to solving this problem is recognizing early that a basin-filling record is commonly
episodic. The concept of cyclothems (Wanless and Weller, 1932) and its many derivatives (for example Kauffman, 1969) recognized recurrent and, therefore,
predictable patterns of deposition in time and space.
Wheeler (1960) and Sloss (1963) developed the concept of
repetitive depositional and erosional patterns within eratonic basins.
The repetitive nature of Cenozoic deposition was recognized early in thick stratigraphic section of the Gulf of
Mexico basin (Deussen and Owen, 1939; Fisher, 1964)
and provided a basis for regional stratigraphic correlation and analysis. Here, successions of sandy tongues
consisting of coastal plain and paralic deposits extend
progressively basinward, where they overlie and grade
into thick marine mud rocks (Figure 1). The thick sandrich offlapping wedges are separated by thinner updipextending tongues of fossiliferous marine mudstone that
effectively punctuate the lithostratigraphy of the upper
5-6 km of the sedimentary section. Sedimentary cycles
were defined by the bounding marine mud-rock units.
These marine-shale bounded units provide the basis for
systematic depositional analysis of the Cenozoic section
of the northern Gulf of Mexico.
A similar stratigraphic patternofflapping sandy
tongues or wedges separated by widespread, often litho-

125

Genetic Stratigraphic Sequences in Basin Analysis I

126
NW

RIO GRANDE EMBAYMENl

O-rO

2-

8 O

"S3.

12

FACES ASSEMBLAGE
CZD Coastal-plain fluvial
E3 Bay/lagoon
E3Paralic- shore-zone/deltaic
I I Shelf and slope
EUD Intraslope basin

520

30km
VE,=40
CONTINENTAL

CRUST
ATTENUATED CONTL.

Figure 1Generalized dip-oriented stratigrapliic cross section tlirougli nortliwestern Gulf Coast sedimentary wedge. Principal
Cenozoic depositional complexes are labeled. Note expansion of complexes across major growth fault zones, which mark positions of successive paleocontinental margins.
logically distinct, marine unitsdominates the depositional pattern in clastic basin fills of diverse ages and
tectonic settings (Figure 2). The marine beds again provide a basis for regional correlation and integrative fades
analysis.
Depositional Elements and Depositional Architecture
Analyses of the depositional history of the northwestern Gulf Coast and late Paleozoic Midland basin fills,
particularly as illustrated in Texas Bureau of Economic
Geology publications (beginning with Fisher and McGowen's [1967] analysis of the Wilcox Group), generally
used transgression-bounded units to define and map
three-dimensional, genetic lithostratigraphic units
termed "depositional systems." Interpreting deposition
systems as fundamental building blocks of basin fills is a
major facet of basin analysis (Miall, 1984; Galloway and
Hobday, 1983).
The increasing quality and availability of regional
reflection seismic sections led to the development of seismic stratigraphy, an approach to basin analysis that
delineates and maps regional depositional and erosional
surfaces (Mitchum et al, 1977). Because disconformable
surfaces are particularly apparent on regional seismic
sections, boundary defined units, or seismic sequences,
become the fundamental element for basin analysis using
seismic data.
Depositional system and seismic stratigraphic
analyses, with their divergent but complementary

emphases on sedimentary volumes and bounding surfaces, incorporate the three key elements that define the
genetic stratigraphy of basin fills.
(1) Depositional systems are three-dimensional assemblages of process-related fades that record major paleogeomorphic basin elements. They grade laterally into
adjacent systems, forming logical associations of paleogeographic elements. The systems commonly display
evolutionary trends through stratigraphic successions
that record geologically significant time spans but are
separated from underlying and overlying systems by
hiatal (disconformable) surfaces. Genetic stratigraphic
packages typically consist of the sediments of several
related depositional systems.
(2) Bounding hiatal surfaces separate stratigraphic
packages and record major interruptions in basin depositional history. These surfaces record significant periods
of nondeposition or very slow deposition, with or without concomitant subaerial or submarine erosion. Surfaces have several origins and may themselves be part of a
migratory fades tract. Thus, hiatal surfaces can be part
of a related time-equivalent depositional system tract or
can separate system tracts of different ages and genetic
stratigraphic units.
Unconformities are hiatal surfaces that demonstrably
truncate underlying strata. Three types of unconformities are recognized: (1) subaerial erosion surfaces, including incised valley systems, (2) shoreface ravinement
surfaces eroded during transgression (Swift, 1968), and
(3) submarine shelf and slope erosion surfaces reflecting
sediment starvation and erosion by currents or mass

William E. Galloway

127

w
j Nonmarine sandslorw and mudstone
x|J Coastal sandstone
I

I Marine mudstone

OFFLAP

gMorine limestone

[Marine mudstone

^'y/|.jCooslol and nonmorine sandstone ond mudstone

Figure 2Generalized stratigraphic cross sections of offlapping basin-margin sedimentary prisms showing repetitive pattern of
progradational sandy tongues, which consist of coastal and nonmarine facies, separated by transgressive marine mudstone or
limestone units. (A) Upper Cretaceous fill of western North American seaway, San Juan basin. Seven episodes of sediment influx
into this intermittently thrust-loaded foreland basin are clearly evident. (B) Late Pennsylvanian prograding Eastern shelf. Midland
basin. Here, eight prominent but thin sedimentary cycles punctuate mixed siliciclastic and carbonate fill of stable intracratonic
basin. Sections modified from Molenaar (1983) and Brown et al (1973).
wasting (Frazier, 1974; Christie-Blick et al, in press).
Condensed sections are the product of very slow deposition. In terrigenous clastic basin fills, marine condensed sections form on the open shelf and slope during
extensive basin-margin transgression and flooding, and
exhibit a variety of paleontologic or compositional
attributes. Thin, widespread, highly fossiliferous hemipelagic and pelagic mudstone drapes commonly reflect
sediment starvation. Chemical sedimentsthin marl or
limestone beds, glauconite, phosphatic zones, siUceous
shaleindicate extremely slow deposition. Widespread
radioactive marine mudstone units (hot shales) similarly
reflect slow sedimentation and concentration of organic
matter. In subaerial environments, widespread paleosoils
and coaly zones indicate slow rates of clastic accumulation.
(3) Bedding architecture describes the geometric relationship between bedding surfaces or the stratification
within depositional systems and at bounding surfaces.
The contrasting geometries of progradational, aggradational, and retrogradational sedimentary units have long
been recognized. Similarly, the hierarchy of erosional

features, ranging from simple channeling to large-scale


valley or canyon incision, has been recognized using outcrop and conventional subsurface data. The ability of
seismic data to resolve surfaces and the geometry of discordant stratification within the framework of those surfaces has increased our understanding and ability to use
bedding relationships in interpreting depositional process and history (Mitchum et al, 1977).
A thorough analysis of a sedimentary basin fill must
incorporate and reconcile the three-dimensional distribution of depositional systems and their component facies,
bedding geometries, bounding surfaces, and condensed
sections within the motif of recurrent depositional cycles
shown in Figures 1 and 2. Sequence stratigraphythe
analysis of repetitive genetically related depositional
units bounded in part by surfaces of nondeposition or
erosionattempts this integration. The genetic stratigraphic sequence I propose emphasizes the equal importance of depositional systems and bounding hiatal
surfaces as elements of the basin fill. The primary objective is to define an operational stratigraphic unit that (1)
groups all sediments that record a common paleogeo-

128

Genetic Stratigraphic Sequences in Basin Analysis I

graphic^ assemblage of depositional systems and (2) is


bounded by stratal surfaces that reflect major reorganizations in basin paleogeographic framework. Given a
choice of possible surfaces, those that separate major
shifts or changes in depositional system organization are
emphasized. Delineation, mapping, and interpretation
of these sequences then provides an overview of the principal depositional episodes recorded within the sedimentary fill of a basin.
Depositional Episodes and Genetic Stratigraphic Sequences

Frazier (1974) developed a conceptual model for defining genetic stratigraphic units and their components in
offlap-filled clastic basins. Using three-dimensional
stratigraphic studies of the late Quaternary depositional
sequences of the northern Gulf Coast basin, Frazier distilled several principles that form a foundation for
sequence stratigraphy (Galloway and Hobday, 1983).
(1) Terrigenous clastic sediments are allochthonous
and must be transported to the basin margin primarily by
fluvial systems. Therefore, major reorganization of
basin paleogeography typically involves changes in principal fluvial axes.
(2) Basins are filled through a repetitive alternation of
depositional (offlap) and nondepositional (transgressive)
intervals. At any specific time, active deposition is concentrated within a small portion of the total submerged
basin area. Minor amounts of terrigenous sediment accumulate elsewhere; nondeposition or erosion may, in fact,
dominate. Consequently, essentially nondepositional
interludes separate depositional intervals.
(3) The time interval represented by resuhant submarine hiatal surfaces varies areally; however, such surfaces
everywhere separate sediments of different depositional
events and, therefore, ages.
(4) Each depositional pulse or event is separated from
other pulses by hiatal surfaces basinward of initial and
terminal shorelines of maximum flooding. The pulse
produces a surface-bounded genetic stratigraphic unit
that Frazier termed "facies sequence." Analogous stratigraphic units have also been called "parasequences"
(Van Wagoner etal, 1987).
(5) Progradational, aggradational, and transgressive
facies are arranged predictably within a facies sequence.
(6) A hierarchy of progradational-transgressive cycles
exists in most basins. Multiple events punctuate regional
depositional episodes.
Frazier recognized depositional episodes and depositional complexes as the principal genetic time and rock
stratigraphic subdivisions of basin history and fill. Depositional episodes are ended by regional flooding events;
their physical stratigraphic units provide a record of
coastal outbuilding capped by transgressive facies and
superjacent submarine unconformities or condensed sedimentary veneers. Although Frazier did not use the term
^As used here, paleogeography emphasizes the areal distribution of principal physical geographic elements of the basin, including major fluvial axes and
related shore-zone and submarine features.

"depositional sequence," his depositional complex is a


sequence-stratigraphic unit bounded by surfaces of erosion or nondeposition and their correlative conformities.
The complex provides the basis for defining an alternative sequence-stratigraphic unit that does not rely on the
presence or recognition of widespread subaerial unconformities, as required in the definition of depositional
sequences proposed by the Exxon research group (Vail,
1987; Van Wagoner et al, 1987). I build upon Frazier's
model to provide an alternative sequence-stratigraphic
paradigm proven useful in analyzing prograding clastic
basin fills.
DEPOSITIONAL EPISODES AND GENETIC
STRATIGRAPHIC SEQUENCES
Progradation of a basin-margin sedimentary prism
during a depositional episode requires that sediment
accumulate in depositional systems ranging from deepwater slope and basin plain to paraUc (deltaic, shore
zone, and shelf) and terrestrial (fluvial or alluvial fan).
Four bathymetric and depositional regimesslope, shelf
edge, shelf, and coastal plainsuccessively pass a reference point as basin-margin progradation proceeds
beyond that point. Shelf-edge progradation, as defined
by the break in depositional slope that separates shelfplatform sediments deposited by traction currents from
slope sediments deposited dominantly by gravitational
resedimentation, is the most stable guide to the extent of
basin-margin outbuilding at any particular time or stratigraphic level (Winker, 1982; Jackson and Galloway,
1984). The shore zone, in contrast, commonly shifts tens
of kilometers across the depositional platform in
response to minor base-level changes or variation in sediment supply. An ideal depositional episode thus is
recorded by a cycle shoreline advance and retreat and a
pulse of shelf-edge progradation and foundering.
Elements of Depositional Episodes and Sequences

The temporal framework and facies stratigraphy of a


genetic stratigraphic sequence produced by an idealized
depositional episode are shown in Figure 3. The timespace diagram at the top of the figure illustrates the temporal and spatial relationships of principal
environmental assemblages. The lower cross section
illustrates the bedding architecture of the genetic stratigraphic sequence. The episode and sequence consist of
three families of elements: offlap components, onlap or
transgressive components, and bounding surfaces
reflecting maximum marine flooding (hiatal surfaces of
Frazier, 1974).
Offlap components (Figure 3) include (1) commonly
sandy fluvial, delta-plain, and bay/lagoon facies that
reflect aggradation of the coastal plain, (2) progradational deposits of the shore zone (also sandy) that overlie,
landward, the flooded depositional platform of the preceding depositional sequence and, seaward, the contemporaneous facies of the offlapping continental slope, and

Willlam E. Galloway

TRANSGRESSIVE
REWORKING,

129

HIATUS^
CHEMICAL SEDIMENTS
CONDENSED SECTION
SUBMARINE UNCONF
PELAGIC DRAPE
SLOPE

RE6RADING

HIATUc

DEPOSITIONAL PLATFORM

SHELF EDGE

SLOPE

Figure 3Idealized stratigrapliic architecture of simple depositional episode and resultant genetic stratigraphic sequence. Upper
diagram (episode) has time as vertical axis. Lower diagram (sequence) shows resultant stratigraphic architecture and facies associations with depth as vertical axis. From Galloway (1987).
(3) mixed aggradational lower slope and progradational
upper slope facies. Although the terms "slope of flap"
and "progradation" are used synonymously, the internal
facies architecture of slope deposits is dominated by
gravitational remobilization and aggradational stacking
of sediment at the toe and on the adjacent basin floor
(Mitchum, 1985; Mutti, 1985). Thus, offlap slope systems more commonly contain a mix of mainly upper
slope progradational facies, which display distinctly
clinoform bedding geometry, and base-of-slope and
adjacent basin-plain aggradational facies sequences,
which display more complex onlap and mounded bedding geometries (Mitchum, 1985). Mutti's (1985) stage
III turbidite deposits are especially common, but stages I
and II sandy base-of-slope lobes also occur interspersed
with the muddier offlap apron. The degree to which
organized submarine fans or disorganized slope aprons
form is a function of several variables, including sediment supply (rate, degree of focusing, texture), basin
hydrography, tectonism, and base-level instability
(Mutti, 1985; Stow et al, 1985). In thick slope successions, gravity tectonics typically modify simply patterns
of facies and stratal architecture by focusing subsidence
at the shelf margin and creating major stratal discontinuity surfaces (Jackson and Galloway, 1984).
Onlap components (Figure 3) consist of (1) reworked
shore-zone and shelf facies deposited during and soon

after shoreline retreat, and (2) an apron of gravitationally resedimented, upper slope and shelf-margin deposits
at the toe of the slope. The transgressive period, following active outbuilding of the continental margin, is a
good time for extensive mass wasting and retrogradation
of the upper slope and continental margin (Dietz, 1963;
Brown and Fisher, 1980; Winker, 1984). As a result, a distinct apron of resedimented material onlaps the toe of the
slope clinoform (Figure 3). Regraded slope aprons may
also consist of Mutti's (1985) types I and II turbidite complexes as well as more localized slump debris. For simplicity, Figure 3 illustrates an idealized genetic
stratigraphic sequence in which little sediment is added
during the transgressive part of the cycle. In this circumstance, the depositional record of transgression may consist of a discontinuous veneer of reworked shore-zone
sediments capping a ravinement surface. Active sediment
input commonly continues during transgression (depositional transgression or common onlap of Curray, 1964),
resulting in thick deposits and recording an extended period of landward-stepping depositional events (Figure
4A). The term "retrogradation" is useful for describing
these long-term periods of shoreline and shelf-edge
retreat.
Finally, the genetic stratigraphic sequence is bounded
by two stratigraphic surfaces (Figure 3) that record the
relative clastic-sediment starvation of the shelf and slope

Genetic Stratigraphic Sequences in Basin Analysis I

130

Depositional |
episode
j^

Continental

margin"

_StiorelineI
-Bosinward

Greolesf marine
influence

Greotesf fluviol
influence

Platform

Extensive low
coostalpioin flooding

deltas

Stielf-edge

deltas

Stielf
Streomploin-

yjf'PX

Aggradational
V ^
faorrier

Tronsgressive barrier
Lagoon

Lagoon

Dcminantly
subaerio!
coastal plain

Figure 4(A) Schematic time-space diagrams illustrating complex history of outbuilding and retrogradation typical of genetic
depositional episodes. Smaller cycles of progradation and retreat punctuate long-term history of regional offlap followed by progressive shoreline retreat. Patterns are same as in Figure 3. (B, C) Location of common genetic facies associations deposited in
deltaic headland and interdeltaic bight transects during depositional episode.

William E. Galloway
during transgression and the ensuing period of maximum
marine flooding. The concept of hiatal surfaces separating the transgressive or retrogradational deposits of one
sequence from the progradational deposits of the succeeding sequence is used here in Frazier's (1974) original
sense. Intermittent sediment deposition commonly
occurs across areas of the flooded depositional platform.
However, terrigenous sedimentation rates are extremely
low, and extended time intervals are recorded by thin,
stratigraphically and compositionally distinctive marker
beds, which contain multiple disconformities that cannot
be resolved individually. Stratigraphic manifestations of
marine-flooding surfaces include both unconformities
and condensed sections, which form downlap seismic
sequence boundaries (Asquith, 1970; Mitchum et al,
1977). The manifestations provide prominent log markers and commonly are petroleum source rock horizons
(Meyer and Nederlof, 1984).
Significantly, Frazier's model also predicts the presence of a subaerial hiatal surface along and within the
landward margin of the genetic stratigraphic sequence
(Figure 3). As the shoreline, which is the focus of deposition, shifts basinward, the inner coastal plain may
become a graded surface that acts as a sediment bypass
zone. Minor changes in base level, or the development of
peripheral uplift that occurs in larger basins as deposition
loads the crust, leads to nondeposition, valley incision,
or even low-angle truncation of older basin-fringing
parts of the sequence. This erosion surface is the key type
1 depositional sequence boundary of Vail et al (1984).
However, its significance as a synchronous regional
stratigraphic boundary is based on the assumption that
stratal architecture of basin margins is uniquely dominated by eustatic falls of sea level to or below the shelf
edge (an assumption that many authors, such as Pitman
and Golovchenko [1983], Miall [1986], and ChristieBlick et al [in press] question). Most recently, Vail (1987)
and Van Wagoner et al (1987) have redefined the type 1
subaerial unconformity, stating it requires a drop of sea
level only below the depositional shoreline break or
shoreface. This redefinition significantly reduces the
amount of sea level fall required to generate a type 1
sequence boundary. At the same time, such redefinition
reduces the potential geographic extent and stratigraphic
significance of, and proportionally increases the tectonic
influence on, the resultant subaerial erosion surface.
All sequences are three-dimensional stratigraphic units
with regional continuity along strike within a basin. Ideally, each sequence encapsulates several related depositional systems in both the dip and strike directions. Both
modern and ancient coastal plains reveal a common pattern of laterally associated depositional elements in many
divergent-margin and foreland basins (Weimer, 1970;
Winker, 1979,1984; Galloway, 1981). Large extrabasinal
fluvial systems produce broad coastal-plain alluvial/
deltaic aprons and deltaic headlands that prograde onto
the shelf edge (Figure 5). Between fluvial/deltaic axes,
minor streams produce a stream plain that grades basinward into an interdeltaic coastal bight. Sediment is transported into the bight by local streams and longshore
reworking from the adjacent dehaic headlands. A shelf

131

extends basinward from the shore-zone depositional system of the bight (Figure 5), and the shelf edge lies offshore. The same paleogeographic elements persist during
transgression and marine flooding; however, a broad
shelf fronts both the deltaic headlands and adjacent
bights. Two different generalized system transects are
required to characterize the sequence stratigraphy of a
depositional episode along such a typical basin margin.
Figure 4B and C illustrates the temporal and spatial
relationships of typical facies assemblages along transects through a deltaic headland and interdeltaic bight.
Sediment supply remains active during coastal retreat,
and onlap is dominated by gradual retrogradation. In the
deltaic headland, initial delta progradation first reclaims
the flooded platform of the previous episode. Prodelta
and delta-front facies comprise the progradational facies
elements. Delta-plain deposits cap the progradational
platform. As progradation extends to the underlying
shelf edge, thickness of shelf-edge, delta-front and prodelta facies increases and gravity remobilization assumes
a more important role in further outbuilding the deeper
water slope (Suter and Berryhill, 1985; Armentrout,
1987). The outbuilding slope includes both progradational elements of the delta system and mixed architectures of an offlapping submarine fan or apron system.
Concomitantly, the inland delta plain continues to
aggrade, ultimately forming a broad aggradational alluvial apron. Nodal avulsion of trunk streams across this
apron results in delta-lobe switching, often at a scale of
hundreds of miles in large systems. (Galloway [1981]
summarized numerous examples from the Quaternary
and Cenozoic record of the Gulf Coast.) Thus, in any one
transect, outbuilding is punctuated by subregional (tens
to hundreds of kilometers along depositional strike)
transgressions which deposit delta-destructional facies.
Retrogradation results in increasingly restricted deltafront and prodelta deposition, and deltas once more prograde into shoal water of the newly flooded platform.
More closely balanced supply and reworking result in
stacked sequences of increasingly marine-dominated
delta-front facies (Galloway, 1975). Delta-destructional
facies, in turn, are covered by the widespread prodelta/
shelf mud blanket. Relative rise in base level also
enhances the aggradation of delta and fluvial plains.
Increased preservation of overbank and flood-basin
facies results (Galloway et al, 1986). The unstable shelfmargin delta deposits continue to slump and resediment
delta-front, prodelta, and upper-slope deposits, creating
an onlapping slope apron. Large submarine canyons may
be incised across the expanding shelf (Galloway et al,
1988).
In interdeltaic areas (Figure 4C), a sandy or muddy
strand plain progrades across the flooded shelf platform
of the previous episode. Subaerial coastal plain deposits,
commonly deposited by many small streams forming a
stream plain (Galloway, 1981), aggrade upon the progradational foundation of strand-plain deposits. Rate of
outbuilding slows at the shelf margin, where muddy sediments extend from the shelf onto the slope. Slope offlap
again consists of mixed progradational and aggradational resedimented deposits. The shelf-slope break is

Genetic Stratigraphic Sequences in Basin Analysis I

132

(A)

Extrabasinai

streom

Interdeltaic

3igh1

RewO'k ng

Deltaic

Headland

Geomorphic or structurol focus

Erosional edge of older depositionol sequence

Figure 5Idealized paleogeographic elements of prograding clastic coastal plain and shelf. One or more fluvial/deltaic headlands
prograde to shelf edge and onto upper continental slope. Shoreline and shelf of interdeltaic bight are fed both by longshore transport of sand and mud and by local streams. Entry of major and secondary fluvial systems onto depositional coastal plain is focused
through erosional valleys at uplifted margin of depositional basin. Nodal avulsion then creates broad alluvial aprons. From Galloway et al (1986).

more gradual than at the more actively prograding deltaic headland; consequent regrading by mass wasting is
probably less intense during the subsequent retrogradation. Because sediment is supplied alongshore from adjacent deltaic headlands with their shifting depocenters,
outbuilding of the interdeltaic facies tract is similarly
sporadic, with bursts of progradation punctuated subregionally by stability or transgression. Retrogradation of
wave-dominated coasts, such as the northwestern Gulf of
Mexico, commonly is reflected by the increasing importance of barrier bar/lagoon systems as coastal flooding
occurs. Rising base level preserves aggrading coastalplain fluvial and bay/lagoon systems (Galloway et al,
1986). Thick aggradational barrier and thin transgressive
barrier facies are preserved. A muddy shelf blanket covers all or part of the broad, newly flooded depositional
platform.
Distinct shelf-system deposits, including sand-rich

facies, are most likely formed during transgression and


flooding (Swift and Rice, 1984). Because shelf deposits
are derived from reworked transgressed or contemporary
retrogradational deposits, their distribution commonly
reflects the paleogeography of the precursor depositional
episode. These deposits are best included in and mapped
as a facies element of the underlying genetic stratigraphic
sequence.
In summary, the genetic stratigraphic sequence is a
package of sediments recording a significant episode of
basin margin outbuilding and basin filling bounded by
periods of widespread basin-margin flooding (Figure 6).
The depositional veneer or erosion surface defining the
time of maximum marine flooding is commonly the
boundary between the major three-dimensional depositional systems. The common stratigraphic juxtaposition
of typically thin transgressive facies or surfaces and the
condensed paleontologic or sedimentologic veneers

William E. Galloway

133

EXPLANATION
I: :^: ;:-l

Shore-zone facies

Genetic stratigraphic sequence boundary and correlative conformity

'

Depositional sequence bounding unconformity and correlative conformity


Depositional surfaces
Successive coastal depositionol systems

(2)
X,

Genetic stratigraphic sequence

Depositional sequence

Figure 6Comparison of boundaries for (A) Exxon type 1 and (B) Exxon type 2 depositional sequences with those of genetic
stratigraphic sequence. (A) E^on paradigm emphasizes subaerial unconformity and its equivalent stratal surface, which may be
quite prominent where relative base level drops below progradational shelf edge. (B) In contrast, bounding unconformity is
obscure and of limited extent in type 2 sequences where base level does not drop below platform margin. In both examples,
downlapped hiatal surface created by transgression and flooding of coastal plain creates an easily correlated horizon that encapsulates prograded wedge of sandy coastal-plain, shore-zone, and marine-slope sediments.

recording maximum transgressive flooding makes their


interchangeable use for regional stratigraphic correlation
practical in many basin settings. Within the sequence, a
coherent assemblage of depositional systems can be
defined and delineated by mapping framework sandstone distribution. This inferred and observed genetic
coherence of the depositional system framework within
marine flooding-surface bounded sequences is a fundamental distinction from the lowstand unconformity
bounded depositional sequence of Vail et al (1984), which
contains the flooding surface and combines the retrogradational facies of the older system with the progradational facies of the younger (Figure 6). Within the genetic
stratigraphic sequence, evolving patterns in coastal-plain
fluvial, deltaic, interdeltaic shore-zone, shelf, and slope
system deposition may be recognized or predicted.

CAUSES OF DEPOSITIONAL EPISODES


Given the prominence of episodic deposition and
resultant sequences in many basin-margin sedimentary
wedges, the question of cause becomes particularly compelling. Certainly, the interpretation that synchronous,
in-phase, eustatic sea level changes control sedimentary
cycles is popular (Vail et al, 1984). However, many workers (Krumbein and Sloss, 1963; Curray, 1964; Hardenbol
et al, 1981) have emphasized that depositional patterns
reflect the dynamic interplay of three principal factors
(Figure 7): eustatic changes, terrigenous sediment supply,
and basin subsidence rate.
Eustatic changes in sea level directly influence the location of the shoreline and, thus, the site where sediment
flows into the basin. The eustatic component includes

Genetic Stratigraphic Sequences in Basin Analysis I

134

Spreading
Rates

Glacial
Eustasy

Geoidal
Eustosy

EUSTATIC
SEA LEVEL CHANGE

TECTONICS

Source/Bosin
Relief

Tectonic
Eustasy

SUBSIDENCE
RATE

Climate

SEDIMENT
SUPPLY

\ /
SEQUENCE
DEVELOPMENT

Figure 7Variables that influence depositional liistory and sequence stratigrapliy of prograding basin margins. Sequence architecture is product of interplay between eustatic sea level, subsidence rate, and sediment supply.
changes in ocean basin volume (tectono-eustasy), water
volume (glacial eustasy), and geoidal surface (geoidal
eustasy) (Fairbridge, 1961; Morner, 1980).
Terrigenous sediment supply is determined by source
terranes and regional climate. Subregionally, autocyclic
processes play an important role.
Basin subsidence rate (exclusive of depositional loadinduced subsidence) is primarily a product of local and
regional plate-margin and intraplate thermal and stress
regimes. Both absolute rate and changes in magnitude or
polarity of rate determine the stratigraphic architecture
of basin margins (Pitman, 1978).
Under suitable conditions, the rates of subsidence,
uplift, sediment accumulation, and eustatic base-level
change can approach maximum but comparable ranges
of 1 to 100 m/l,000 years (Figure 8). Thus, as shown in
Figure 9, a full range of depositional architectures
reflecting basin-margin progradation, aggradation, retrogradation, and transgression can result equally from
variations of sediment influx, subsidence rate, or eustatic
sea level. In basic architecture, the stratigraphic product
looks similar regardless of which variable is changed.
Genetic stratigraphic sequences are simply combinations
of progradational components followed by retrogradational or transgressive components. Aggradational intervals may also be incorporated at various levels in the
sequence. Therefore, all three factorssediment supply,
subsidence rate, and eustatic sea levelmust be considered in any critical examination of basin-margin stratigraphic evolution (Miall, 1986).

Eustatic Sea Level Change


Eustatic sea level has three components of potential
change: glacial, tectonic, and geoidal (Figure 7).
Glacial eustatic changes result from the changing volume of continental ice caps and have potentially high
change rates of 10 to 100 m/l ,000 years (Pitman, 1978).
Such changes are limited but not inherent to periods of

large-scale polar ice cap formation. Polar ice caps began


to form during the Oligocene and were widespread by the
Miocene (Loutit and Kennett, 1981; Leckie and Webb,
1983). In contrast, the later Mesozoic and early Cenozoic
world was probably largely ice free. Large volumes of ice
are required to affect sea level by tens of meters. For
example, a sea level rise of about 60 m would result if all
present glaciers and ice caps melted (Pitman and Golovchenko, 1983). Lack of direct geologic evidence for comparably large volumes of ice during the Mesozoic and
early Cenozoic is difficult to reconcile with theorized
high-amplitude glacial eustasy during these times. We
presume glacial eustasy is an important sea level control
during parts of earth history, but it remains an unlikely
mechanism for other, equally extended periods.
Tectono-eustatic changes result from changes in ocean
volume by large-scale lithospheric plate interactions. The
most important of these appears to be variation in rate of
sea-floor spreading and consequent elevation of spreading ridges and surrounding oceanic crust. Pitman (1978)
calculated maximum possible rates of sea level change
due to changing ridge volume to approximate 1 cm/1,000
years. A general lowering of sea level by 100 to 300 m has
occurred due to decUning spreading rates since the Late
Cretaceous (Kominz, 1984), but tectono-eustasy seems
incapable of explaining rapid (100,000 years), stratigraphically substantive (10-100 m) changes in sea level.
Less appreciated by stratigraphers is the potential role
of geoidal eustasy. The geoid is the equipotential surface
of the combined rotational and gravitational potential
fields and corresponds to mean geodetic sea level. The
ambient geoid (sea level surface) commonly varies from
an ideal mathematical surface by several tens of meters.
Changes in geoid geometry at scales comparable to other
crustal tectonic processes would be expected. However,
Morner (1980) predicted recurrent, rapid (1,000 years)
variations in geoid elevation of magnitudes of tens of
meters. Newman et al (1980) suggested evidence for such
changes in historical geodetic data. Aside from providing
a speculative mechanism for nonglacial rapid eustatic sea
level change, Morner's (1980) hypothesis, if validated.

William E. Galloway

135

o
a

100.0

u>
c

10

lOOO.Or

c
a>

a "Ji
o

to

O
(U
o
w

c
o

0)

O
O
O

J3

10.0

o
CD

CO

(U

.=

<i>

x:

_
^

1.0
0.1

0)

3
O
tf>

i
o

o
O

o
o

c O
<u
o>
o
o
c
o
_
-

a.
UJ

V)

o
4)

a
E

n
O

<i>

1
1
1

o
c
o
o

0-.

1
1
0-.

I
a

a.

0.01
0.001

ACCUMULATION

DENUDATION

UPLIFT-

SUBSIDENCE

EUSTASY

Figure 8Comparative rates of deposition, erosion, tectonic uplift or subsidence, and eustatic sea level change. Rates exceeding 1
m/1,000 years characterize a broad range of depositional and tectonic settings. Only glacial and possibly geoidal eustasy match
maximum uplift/subsidence and accumulation rates. Modified from Stow et al (1985).

has important implications for sequence stratigraphy.


First, geoidal eustatic changes, although synchronous
and global, would show basin-to-basin phase shifts. Rise
in sea level in one area must be balanced by fall in other
areas. Second, the data for ongoing geoidal change
(Newman et al, 1980, p. 555) "...paradoxically demonstrate the futility of searching for a purely eustatic curve.
The geoid is continually deforming....''

Transgressive

Retrogradational
Aggradational

.*-Progradational

Figure 9Schematic stratigraphic architectures of progradation, aggradation, retrogradation, and transgression with three
variables that control their formation. Change in any of three
variables potentially yields similar stratigraphic architecture.
Genetic stratigraphic sequence typically includes superposition
of successive architectures, reflecting evolution from shoreline
and shelf-margin progradation to shoreline retreat and shelfmargin foundering. From Galloway (1987).

Subsidence Rate
Subsidence is a product of tectonics (crustal extension,
cooling, tectonic loading) or sedimentary loading (Figure
7). Most subsidence models assume uniform rates that
can be expressed by simple mathematical formulas.
However, uniform subsidence is only a first-order
approximation of reality; cooling, stretching, or loading
rates need not be uniform (Wiltschko and Dorr, 1983;
Poag and Schlee, 1984; Ru and Pigott, 1986). Further,
loading and flexural downbowing of the crust induce
stress that results in peripheral uplift. The locus of such
uplift is migratory if loading varies in time or space
(Quinland and Beaumont, 1984). Changes in regional
intraplate stress fields significantly modify the crust's
flexural response to loading. Models demonstrate that
changes of a few kilobars in the horizontal stress regime
can induce relative base-level changes by 100 m
(Cloetingh et al, 1985).
Evidence suggests that magnitude and orientation of
horizontal crustal stress fields can change over periods 1
to 10 m.y.; therefore, base-level changes of 1 cm/1,000
years are possible (Cloetingh et al, 1985) in basins next to
active tectonic zones. The intracontinental scale of
crustal stress regimes (Zoback and Zoback, 1980) indicates the potential for changing plate-margin tectonism
to influence subsidence across broad areas of continental
plates.
Sediment Supply
Sediment supply has been neglected in sequence
analysis of basins not directly associated with convergent

136

Genetic Stratigraphic Sequences in Basin Analysis I

plate margins. However, sediment supply is important in


prograding basin margins where input inherently must
exceed the accommodation potential of the receiving
basin. Observed Quaternary depocenter sedimentation
rates commonly are tens of meters per thousand years,
matching or exceeding all but the most rapid glacial
eustatic sea level change rates (Figure 8). Schumm (1977)
discussed factors affecting sediment supply. Overall terrigenous clastic sediment supply is a complex response to
source-to-basin relief and climate of the source area (Figure 7). Relief, in turn, is controlled by tectonics in the
source terrane and the intermediate transport pathways
to the basin (which can divert or pirate trunk streams and
create sediment sinks). Changing climate, not climatic
stability, results in surges of sediment supply.
Sediment supply as a major variable is well documented by Stott (1984) for the northern Cretaceous seaway, where average sediment accumulation rates for the
basin varied repeatedly by a factor of two. Swift and Rice
(1984), among other workers, related variations in the
seaway sediment supply to periods of active thrusting in
western North America, establishing a direct connection
between source area tectonism and episodic progradation of the basin margin. Preliminary results of an examination of regional Cenozoic depositional rates in the
northwestern Gulf of Mexico indicate up to tenfold
changes over a few million years. Paleogeographic mapping in both the Cretaceous seaway and Gulf basin has
shown repeated geographic shifts in deltaic depocenters.
Thus, local sediment should show even greater variation
with time.
Sea level change, however, has little effect on quantity
or nature of sediment supply. A fall in sea level would not
increase the volume or caliber of stream sediment load
unless the ratio of drainage basin relief (elevation of the
bedrock source terrane) to transport path length were significantly increased (Schumm, 1977; Vita-Finzi, 1986).
In fact, the opposite is likely to result; at lowstand, the
low-relief coastal plain will be widest and the average
stream gradient least. Volumes of sediment excavated by
valley incision of coastal plains exposed at relative lowstands are inconsequential in the context of sequence sedimentary volumes. Basins cannot be filled by
self-cannabilization.
Comparative Influence of Three Variables
The packaging of terrigenous clastic basin fill into a
series of sequences is the consequence of the ongoing
interplay of three interacting but independent variables.
As in any natural system, one variable may dominate.
Commonly, however, two or even all three variables create the sequence stratigraphic framework of basin margins. Such a three-component system can be readily
represented on a ternary diagram such as shown in Figure
10. If basin analysis is to live up to its potential in elucidating the more subtle tectonic history of basins and their
source terranes and in testing concepts of global eustasy
(including phase-shifted geoidal eustasy), genetic
sequences must be placed appropriately in such a process

Figure 10Ternary diagram with principal factors that control


history of basin infilling as end members. Arrows indicate
trends or attributes that can help classify depositional episodes
and resultant stratigraphic sequences as supply, subsidence, or
eustatic dominated. Most real sequences plot within triangle
rather than as pure end members.
framework. Criteria for distinguishing the effects of subsidence, sediment supply, and eustasy must be developed,
refined, and rigorously applied.
The process triangle shown in Figure 10 illustrates
some trends that may help distinguish one or more end
members.
(1) Thick sedimentary packages indicating high sedimentation rate suggest sediment input as a major control
on sequence organization. Changes of a few tens of
meters in base level logically have little influence on the
long-term paleogeography of depositional settings, such
as major deltaic depocenters, where depositional topography is commonly tens to hundreds of meters and is sustained by complex autocyclic processes of progradation
and destruction as well as load and compaction-induced
subsidence.
(2) Abrupt juxtaposition of genetic facies that reflect
very different parts of the facies tract suggested eustatic
or tectonosubsidence control of sedimentation. Examples include valley incision directly into marine or prodelta facies, downward shift in coastal offlap to a
position below the depositional shelf edge or shoreline
break, and abrupt superposition of shallow-water sediments and faunas on deep-water sediments (Weimer,
1984; Vail et al, 1984). However, care must be exercised.
Retrogradational facies tracts may simply reflect waning
sediment supply in the face of ongoing subsidence.
Transgression, with superposition of marine facies
directly upon nonmarine deposits, may result from rapid
relative sea level rise or from abrupt truncation of sediment supply. In areas of high marine energy, little sediment may accumulate in the inner to middle shelf; the
resultant progradational stratigraphic record would
superimpose shoreface sediments directly on outer
neritic facies.
(3) Downward shifts in coastal onlap require a relative
sea level fall. However, independent evidence must con-

William E. Galloway
firm the coastal depositional facies of the onlapping
wedge. We must exercise care to avoid misinterpreting
geometrically similar gravity glide wedges and
megaslumps (Galloway, 1986).
(4) Abrupt changes in the average depositional rate,
grain size (caliber), or mineralogic composition of sediments comprising successive sequences suggest supply is
a principal control. Changes in one or more of these variables likely reflect regionally significant tectonic or climatic reorganization of source terranes and transport
pathways. Conversely, sea level exerts no direct influence
on the composition and little direct influence on rate of
sediment supply to the basin.
(5) Widespread interplate correlation and demonstrable synchroneity of sequences and their bounding hiatal
surfaces support eustatic sea level change as the principal
control. Again, we must be cautious; correlation and
demonstration of synchroneity must be rigorously
tested. Large-scale tectonic events commonly affect several adjacent plates and may approach global dimensions
(Bally, 1980; Schwan, 1980). Consequently, basin-wide
or even interregional correlation does not rule out tectonics as the cause of sequence development. A hierarchy of
sequence correlations needs to be applied and specified:
Correlation is restricted to one depocenter or to
closely related depocenters.
Correlation extends along an entire plate margin.
Correlation extends between adjacent crustal
plates.
Correlation is global.
The test of synchroneity needs to be carefully appUed.
As illustrated in a series of experiments by Zeller (1964),
the assumption that correlation exists ensures that a correlation will be found, regardless of the reality of the
assumed genetic association. If correlation is presumed,
the complexities and incompleteness of the stratigraphic
record will almost always permit correlation. In-phase
eustatic sea level control has always been an appealing
concept because it brings order to an apparently chaotic
group of strata in different basins and places the fundamental causal mechanism comfortably within the
domain of the stratigrapher. However, the assumption of
eustatic control too easily leads to correlations in the
absence or even in the face of rigorous documentation.
The interaction of rigid plates on the closed surface of a
sphere ensures that tectonic events inherently will be
interregional and approximately synchronous. Tectonism and its sedimentary response commonly will show
slight time shifts from place to place (see Hubbard,
1988). Careful documentation of temporal correlation of
depositional episodes, rather than defaulting to a visual
best fit to presumed global standards, is required if stratigraphy is to contribute to the broader quantitative
understanding of the earth's tectonic and eustatic history.
(6) Presence of intrabasinal angular unconformities,
syndepositional fauUs, or other evidence of contemporaneous tectonism supports subsidence history as a principal control of depositional episodes. Rates of uplift or
subsidence at transform and convergent plate margins
are comparable to maximum rates of eustatic change

137

(Figure 8) and can be expected to strongly influence or


dominate the stratigraphic record in some settings.
(7) Progressive bathymetric deepening over periods of
100,000 years suggests that variable but rapid subsidence
controls patterns of deposition.
Most depositional episodes are the product of a
dynamic interplay between subsidence, sediment supply,
and eustatic change; resultant sequences, thus, would
plot somewhere within the process triangle rather than at
an apex. A major goal of sequence analysis is to determine where individual sequences lie. Additional criteria
for such interpretation need to be developed and tested.

CONTRASTING APPROACHES TO
SEQUENCE ANALYSIS
Many similarities exist between the genetic stratigraphic sequence as described here and the depositional
sequence defined and described by Vail et al (1984) and
more recently by Van Wagoner et al (1987). The genetic
stratigraphic sequence is most analogous to the parasequence set of Van Wagoner et al (1987). All have a common conceptual origin in the work on depositional
episodes and depositional complexes of Frazier (1974).
However, emphasis on selection of key bounding surfaces and interpretive goals diverged.
The Exxon group focused on using seismic data and
developing seismic stratigraphy. Seismic reflections
inherently delineate the distribution and geometry of
interfacesdepositional or erosional surfaces. A stratigraphy of surfaces emerges. Depositional systems and system tracts are interpreted by their stratigraphic and
geographic relationship to sequence-bounding unconformities. In contrast, the application of threedimensional facies analysis using subsurface data and
emergence of the depositional systems concept subordinates stratal surfaces to the depositional facies of basin
fills. The genetic stratigraphic sequence, as proposed
here, synthesizes surface stratigraphy within the threedimensional facies framework of depositional systems.
The sequence retains the sedimentologist's emphasis on
interpreting environments by internal features and facies
geometries.
Several key differences distinguish the genetic stratigraphic sequence from the seismically based depositional
sequence, particularly because the latter has been primarily interpreted to reflect ubiquitous eustatic control of
basin filling.
(1) Although the more general definition of Van Wagoner et al (1987) excludes specific reference to sea level
control, their discussion and the conclusion expressed by
Vail (1987) in a companion paper that a sequence is interpreted to be deposited during a cycle of eustatic change of
sea level leaves no doubt that the concept remains faithfully wedded to a stratigraphic framework dominated by
eustasy. A consequence of this emphasis is that many key
elements of the Exxon sequence model, such as widespread, synchronous, subaerial unconformities and
stratigraphically disjointed highstand and lowstand

138

Genetic Stratigraphic Sequences in Basin Analysis I

wedges (Figure 6A), are closely tied to eustatic domination of sediment supply and subsidence. The genetic
stratigraphic sequence paradigm preserves and emphasizes Frazier's (1974) conclusion that a depositional
sequence is a complex of facies derived from common
sources along the basin margin and deposited in a period
of relative base level or tectonic stability. This paradigm
more flexibly accommodates the possible domination by
any of the three variables controlling depositional episodes.
(2) As a consequence of this evolution toward contrasting inferred relationships between major sedimentary
pulses and relative base level stability, proposed genetic
stratigraphic sequence and depositional sequence boundaries are 180 out of phase (Figure 6). Vail (1987) equated
progradation of the basin margin to periods of sea level
fall and lowstand. One result of lowstand is development
of a widespread subaerial unconformity (either type 1 or
2) that serves as the sequence boundary. The stratigraphic
sequence incorporates the pulses of sediment input and
basin-margin progradation (regardless of cause) as the
principal depositional episode and separates these episodes by periods of transgression and flooding. Simply
stated, the Exxon depositional sequences are centered on
marine-flooding events; the genetic sequence is bounded
by them (Figure 6).
(3) The two sequence models place different emphases
on the timing, process, and role of shelf-margin erosion
and retrogradation. In the Exxon model, rapid sea level
fall below the shelf edge results in incised subaerial valleys, upper slope bypass and erosion, and deposition of
lowstand submarine fans (Vail et al, 1984; Mitchum,
1985). Slowing of fall and sea level stability lead to canyon filling. In the stratigraphic sequence model, shelfedge and slope erosion and retrogradation are ongoing
processes controlled by the inherent instability of the
shelf edge and upper slope (a function of rate of progradation, tectonism, and extent of gravity tectonics), and
by the temporal and paleogeographic variations in rate of
sediment supply, basin hydrography, coastal and shelf
morphology, and base-level change. Submarine canyon
excavation and filling as well as fan deposition can occur
at various times within a depositional episode. Deposition of predictable marine onlap wedge (Figure 3) and
excavation of the largest submarine canyons often
accompany initial transgressive flooding of an actively
prograded shelf margin (Galloway et al, 1988).

Strengths of Genetic Stratigraphic Sequence Model


Sequence stratigraphic analysis provides a powerful
tool for basin analysis. However, two fundamentally different sequence models exist. A stratigraphic framework
that recognizes and maps three-dimensional depositional
systems as well as surface-defined genetic sequences will,
I believe, yield a more complete understanding of the
depositional, structural, and eustatic history of sedimentary basins for the following reasons.
(1) The surface or stratigraphic veneer produced by

basin-margin flooding has physical stratigraphic


featureswidespread marine beds or submarine erosion
surfacesthat are easily recognized, correlated, and
mapped using a variety of data bases (ranging from well
logs and seismic data to outcrops) by a broad cross section of earth scientists. Figures 1 and 2 reveal how prominent marine shale or limestone-bounded units (which
contain the surface of maximum flooding) are and why
they have become the basis for conventional stratigraphic and facies nomenclature.
Thin marker beds are often found and can be recognized readily in single outcrop or well sections. The period of maximum flooding produces a condensed
depositional and paleontologic marker. On seismic
records, maximum flooding is recorded by coastal onlap
followed by downlap. Marker beds and hemipelagic
drapes commonly produce high-amplitude continuous
reflectors that can be traced across the depositional platform and down the slope. As noted by Haq et al (1987, p.
1160),' 'The most readily identifiable surface is the transgressive surface...The second most easily recognizable
surface in outcrops is the surface of maximum flooding...." In contrast, subaerial unconformities remain a
most difficuk and subtle target to recognize, particularly
if high-quality regional seismic data sets cannot be
obtained.
(2) More important than ease of recognition, the surface of maximum flooding is a useful stratigraphic
boundary. This surface is well developed across the middle of the facies tract where marine and nonmarine depositional systems interfinger. The surface can be
extrapolated into nonmarine and deep marine sections
(Galloway et al, 1982, 1986). For example, Bouma and
Coleman (1985) used highstand pelagic drapes to subdivide depositional sequences in the Mississippi fan. With a
full range of well, seismic, and paleontologic data available, Armentrout (1987) used the pelagic drapes deposited
during sea level highstands to bound Pleistocene stratigraphic sequences of undoubted eustatic origin that also
contain lowstand unconformities.
(3) The fossil-rich condensed section deposited after
transgression is datable by paleontology and likely incorporates planktonic forms useful in high-resolution
chronostratigraphic correlation. Regardless of sequence
boundary chosen, the marine condensed sections ultimately provide the chronostratigraphic framework for
interregional sequence correlation.
(4) Because maximum flooding and the corresponding
depositional hiatus immediately follow transgression,
the surfaces and sedimentary veneers produced by shoreface reworking and shelf submergence can often be combined as a regional marker horizon to correlate sequence
boundaries and map depositional systems. Nummedal
and Swift (1986) provided an excellent example.
(5) Transgressive and flooding surfaces punctuate the
stratigraphic record. Ubiquitous widespread unconformities associated with abrupt sea level falls are problematical. Their presence depends upon the dominance
of repeated eustatic falls to control basin stratigraphy.
No suitable mechanism for such rapid falls, other than
glacial (and possibly geoidal) eustasy, has been docu-

William E. Galloway

mented. However, the ever-changing interplay of sediment supply, subsidence, and more gradual eustatic
change, which would not necessarily create widespread
terrestrial erosion surfaces, can produce depositional
sequences at any point in time. It is premature to base a
sequence stratigraphic framework upon recognizing
widespread, correlative, subaerial erosion surfaces
which, in turn, are dependent on the dominance of a single variableeustatic sea levelthat is assumed to fluctuate rapidly for unknown causes. Sequences based on
obvious flooding events require no presumption of cause
and can readily incorporate erosion surfaces resulting
from rapid base-level fall, if present.
(6) Genetic stratigraphic sequence boundaries defined
by the maximum flooding surface and correlative facies
of the shoreline of maximum transgression have a geographic extent comparable to Exxon's type 1 depositional sequence boundaries (Figure 6A). They are
difficult to correlate into the nonmarine coastal section.
However, type 1 unconformities are difficult to trace into
the deep marine section where pelagic drapes reflecting
relative highstand are widely used for sequence correlation (Bouma and Coleman, 1985).
The greater extent and correlation potential of maximum flooding surfaces are most obvious where depositional sequences are bounded by type 2 unconformities
(Figure 63). Here, the subtle sequence-defining erosion
surface is found only along the innermost fringe of the
basin fill (in Figure 2, approximately 20% of the facies
tract). In contrast, the alternation of offlap and transgression provides an obvious depositional pattern over
half of the cross section.
The potential for sequence definition by type 2 unconformities is further limited because the most landward
area of basin fill is least likely to be preserved and the
bounding unconformity is easily lost in a maze of fluvial
channelization and local bypass surfaces. At worst, maximum flooding and transgressive surfaces are as extensive as type 1 sequence boundaries; at best, they have
greater lateral extent and stratigraphic visibility than type
2 unconformities.
(7) Transgression and flooding produce robust stratigraphic markers. Basins are inherently dominated by
subsidence. Relative base level can fall only if eustatic sea
level drops at a faster rate than subsidence (or if absolute
uplift of the basin margin occurs). In contrast, relative
rise in base level always accompanies a rise in sea level.
Figure 11 illustrates the results of superimposing typical Eocene sea level fluctuations (Haq et al, 1987) on
three basin-margin segments characterized by slow,
intermediate, and rapid subsidence, respectively. In the
rapidly subsiding basin segment, relative base level
merely undergoes four periods of alternating very rapid
and comparatively slow rise. If sediment supply
remained constant, the stratigraphic record would consist of alternating transgressive and progradational facies
tracts and no subaerial unconformities could be generated. The moderately subsiding basin segment experiences only one brief period of slight relative fall in base
level, perhaps capable of generating a type 2 sequencebounding unconformity. The same four accelerated rela-

139

EXPLANATION
Relatn/e bose-leiiel foil
potentiol for subaerial erosion

Mosf rapid base-lewel rise


maximum potential for transgression

Figure 11Effect of different basin-margin subsidence rates


on depositional response to fluctuating eustatic sea level. Sea
level curve is a 4.5 m.y. segment of proposed Eocene eustatic
curve of Haq et al (1987). Three lower plots show four eustatic
fluctuations superimposed on uniform subsidence curves of 50,
100, and 200 m/m.y. These values represent calculated subsidence rates for updip (near outcrop), mid-dip, and moderately
deep segments of typical Eocene sequences in northwest Gulf
basin. Thus, they represent realistic range of values for basinmargin subsidence in actively filling clastic basins. In margins
characterized by slow subsidence of 50 m/m.y., sea level fall
twice exceeds subsidence rate, creating two falls in relative base
level and two subaerial unconformities. In areas of moderate
subsidence rate (100 m/m.y.), sea level falls reduce effective
subsidence rate, resulting in accelerated progradation. Only
one minor stiUstand or slight rise occurs and single subaerial
erosion surface results. In areas of rapid subsidence, eustatic
changes superimpose a mild oscillatory overprint on base-level
curve dominated by continuous rise. In all examples, times of
sea level rise cause increased relative base-level rise, accompanied by contemporaneous shoreline retreat and marine flooding along entire basin margin. Correlative flooding surfaces are
created throughout basin; unconformities are more localized.

140

Genetic Stratigraphic Sequences in Basin Analysis I

tive base-level rises and marine-flooding events persist.


Subjecting the slowly subsiding basin segment to the
same eustatic cycle results in two relative base-level falls
and subaerial unconformities and the same four relative
rises and associated transgressive fades tracts.
Eustatic sea level rise thus is reflected in all segments by
transgression and creation of synchronous maximumflooding surfaces. Eustatic falls create zero, one, or two
bounding surfaces; only sediment-starved parts of the
basin margin exhibiting very slow subsidence rates would
record all four sea level falls with subaerial unconformities. The full history of eustatic sea level cycles would, of
course, be best preserved in parts of the basin margin in
which subsidence and sedimentation rates are moderate.
Because subsidence varies within and along basin margins, oscillating sea level will create erratic, discontinuous subaerial unconformity surfaces (Parkinson and
Summerhayes, 1985), but the record of sea level rise will
be predictable and will correlate across all coastal and
marginal-marine depositional systems. Generally, transgressive and highstand fades are most widely and easily
correlated in regional basin analysis.
(8) Principal changes in the paleogeographic distribution of depositional systems and depocenters accompany
transgression and flooding events. Major shifts in fluvial
axes, deltaic depocenters, and interdeltaic shore-zone
systems occur between genetic stratigraphic sequences.
Regional mapping of depositional systems and interpreted paleogeography of the succession of major and
minor Cenozoic episodes of the northwest Gulf basin
shows that sequences defined by transgressions record
deposition under conditions of source-area, tectonic,
and base-level stability. Vertical persistence of similar
depositional elements within genetic sequences is, in fact,
a requirement for successful areal delineation of depositional systems from regional lithofacies maps that incorporate hundreds to thousands of meters of sediment (see
Fisher and McGowen, 1967; Galloway et al, 1982). In
contrast, combining the lowstand and transgressive system tract with the overlying highstand system tract, as in
the Exxon model, randomly intermixes deltaic depocenters and interdeltaic systems within the same sequence.
Sedimentary packages bounded by maximum flooding
surfaces are designated genetic stratigraphic sequences.
They consist of temporally and spatially persistent depositional systems and genetically associated fades. Paleogeographic stability does not preclude the presence of
eustatic lowstands and related unconformities. A
eustatic fall of sea level results in entrenchment and consequent stabilization of drainage axes (Fisk, 1944; Suter
and Berryhill, 1985). Deltaic and shore-zone systems
shift basinward, but paleogeography otherwise changes
little. In contrast, transgression and consequent rise in
relative base level aggrade the proximal fluvial systems
(Fisk 1944; Galloway et al, 1986). Thus, periods of transgression and high relative sea level are more likely times
of regional river avulsion, depocenter shifting, and
paleogeographic reorganization. In supply dominated
sequences, the tectonic disturbance creating renewed sediment influx may significantly alter source-area and
trunk-stream morphology, shifting axes of sediment

input and creating a new paleogeographic arrangement


of depositional systems.

DISCUSSION
The depositional episode and its genetic stratigraphic
sequence offer an alternative to the depositional
sequence model, which emphasizes ubiquitous eustatic
control of stratigraphic patterns at all scales. The two
models use many of the same stratigraphic features. Both
models apply to many basin fills; in certain tectonic settings, one or the other may prove more useful. The
genetic stratigraphic sequence paradigm emphasizes
fades associations and depositional systems for defining
fundamental stratigraphic packages. This paradigm further recognizes the equal importance of variable sediment supply and subsidence rates with eustatic sea level
fluctuations. The paradigm should be more readily transferred, with appropriate modification, to the basin fills
of tectonically active plate margins where neither uniform subsidence rate nor sediment supply is a reasonable
assumption. Large-scale first-order and second-order
sequences, created by major episodes of tectonism and
sediment influx and lasting millions of years, in turn,
may be punctuated by eustatic events that determine
details of sediment transport and deposition. We must
recognize a hierarchy of stratigraphic and facies
sequences reflecting combinations of causal mechanisms
to fully understand basin-fill stratigraphy and to better
explore for natural resources.
The flooding-surface bounded sequence model will
need to be expanded to incorporate regional unconformities produced by major tectonic reorganizations of basin
margins (Hubbard et al, 1985). Such expansion is well
within the emphasis of the original Frazier (1974) model.
A complete record of basin filling will lead to two halfsequences: retrogradation or transgression necessarily
accompanies initial marine flooding of a basin; the final
infilling episode deposits only progradational and associated nonmarine aggradational deposits, capped by a subaerial unconformity.
The sedimentary basin fill has been described as the
recorder of geologic history. Analysis of basin fills using
concepts of genetic stratigraphic sequence analysis and
recognizing the complex interplay of tectonic, eustatic,
and internal as well as external sedimentary controls may
resolve major problems in earth science. Stratigraphers
can fine-tune histories of plate interactions and stress/
strain evolution within crustal plates, examine models of
crustal response to tectonic and crustal loading, and perhaps test alternative models of core/mantle interaction
by examining evidence for paleogeoidal deformation.
However, should we become passive servants of a singular simplified framework for interpreting the stratigraphic record, we face the danger of becoming
superfluous to the multidisciplinary family of earth scientists who will advance our understanding of the
dynamic history of the earth.

William E. Galloway
REFERENCES CITED

Armentrout, J. M., 1987, Integration of biostratigraphy and seismic


stratigraphy: Pliocene-Pleistocene, Gulf of Mexico: Gulf Coast
Section of SEPM Eighth Annual Research Conference Proceedings, p. 6-14.
Asquith, D. O., 1970, Depositional topography and major marine environments. Late Cretaceous, Wyoming: AAPG Bulletin, v. 54, p.
1184-1224.
Bally, A. W., 1980, Basins and subsidence-summary, in A. W. Bally, P.
L. Bender, T. R. McGetchin, and R. 1. Walcott, eds.. Dynamics of
plate interiors: Washington, D.C., American Geophysical Union
Geodynamics Series, v. 1, p. 5-20.
Bouma, A. H., and J. M. Coleman, 1985, Mississippi fan: Leg 96 program and principal results, in A. H. Bouma, W. R. Normark, and
N. E. Barnes, eds.. Submarine fans and related turbidite systems:
New York, Springer-Verlag, p. 247-257.
Brown, L. P., Jr., and W. L. Fisher, 1980, Seismic stratigraphic interpretation and petroleum exploration: AAPG Course Notes 16,
181 p.
A. W. Cleaves 11, and A. W. Erxleben, 1973, Pennsylvanian
depositional systems in north-central Texas: University of Texas at
Austin, Bureau of Economic Geology Guidebook 16,122 p.
Christie-Blick, N., G. S. Mountain, and K. G. Miller, in press, Seismic
stratigraphic record of sea-level change, in Studies in geophysics:
Washington, D.C., National Academy of Sciences.
Cloetingh, S., H. McQueen, and K. Lambeck, 1985, On a tectonic
mechanism for regional sea level variations: Earth and Planetary
Science Letters, v. 75, p. 157-166.
Curray, J. R., 1964, Transgressions and regressions, in R. L. Miller,
ed.. Papers in marine geology, Shepard commemorative volume:
New York, Macmillan, p. 175-203.
Deussen, A., andK. D. Owen, 1939, Correlation of surface and subsurface formations in two typical sections of the Gulf Coast of Texas:
AAPG Bulletin, v. 23, p. 1603-1634.
Dietz, R. S., 1963, Wave-base, marine profile of equilibrium, and
wave-built terraces: a critical appraisal: GSA Bulletin, v. 74,
p. 971-990.
Fairbridge, R. W., 1961, Eustatic changes in sea level, in Physics and
chemistry of the earth, v. 4: New York, Pergamon Press, p. 99-185.
Fisher, W. L., 1964, Sedimentary pattern in Eocene cyclic deposits,
northern Gulf Coast region: Kansas Geological Survey Bulletin 169,
p. 151-170.
and J. H. McGowen, 1967, Depositional systems in the Wilcox
Group of Texas and their relationship to occurrence of oil and gas:
Gulf Coast Association of Geological Societies Transactions, v. 17,
p. 105-125.
Fisk, H. N., 1944, Geological investigation of the alluvial valley of the
lower Mississippi River: U.S. Army Corps of Engineers Report,
78 p.
Frazier, D. E., 1974, Depositional episodes: their relationship to the
Quaternary stratigraphic framework in the northwestern portion of
the Gulf basin: University of Texas at Austin, Bureau of Economic
Geology Geological Circular 74-1, 28 p.
Galloway, W. E., 1975, Process framework for describing the morphologic and stratigraphic evolution of deltaic depositional systems, in
M. L. Broussard, ed.. Deltas: Houston, Houston Geological Society, p. 87-98.
1981, Depositional architecture of Cenozoic Gulf Coast plain
fluvial systems: SEPM Special Publication 31, p. 127-155.
1986, Growth faults and fault-related structures of prograding
terrigenous clastic continental margins: Gulf Coast Association of
Geological Societies Transactions, v. 36, p. 121-128.
1987, Depositional and structural architecture of prograding
clastic continental margins: tectonic influence on patterns of basin
filling: Norsk Geologisk Tidsskrift, v. 67, p. 237-251.
and D. K. Hobday, 1983, Terrigenous clastic depositional systems: New York, Springer-Verlag, 423 p.
- W. R. Dingus, and R. Paige, 1988, Depositional framework
and genesis of Wilcox submarine canyon systems, northwest Gulf
Coast: AAPG Bulletin, v. 72, p. 187-188.
D. K. Hobday, and K. Magara, 1982, Frio Formation of the
Texas Gulf Coast basin: depositional systems, structural framework, and hydrocarbon origin, migration, distribution, and exploration potential: University of Texas at Austin, Bureau of
Economic Geology Report of Investigations 122, 78 p.
L. A. Jirik, R. A. Morton, and J. R. DuBar, 1986, Lower Mio-

141

cene (Fleming) depositional episode of the Texas coastal plain and


continental shelf: structural framework, facies and hydrocarbon
resources: University of Texas at Austin, Bureau of Economic
Geology Report of Investigations 150, 50 p.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of nuciuating sea levels since the Triassic: Science, v. 235, p. 1156-1166.
Hardenbol, J., P. R. Vail, and J. Ferrer, 1981, Interpreting paleoenvironments, subsidence history and sea-level changes of passive margins from seismic and biostratigraphy, in R. Blanchert and 1.
Montadert, eds.. Geology of continental margins: International
Geological Congress Proceedings, Oceanologica Acta, v. 4,
p. 33-44.
Hubbard, R. J., 1988, Age and significance of sequence boundaries on
Jurassic and Early Cretaceous rifted continental margins: AAPG
Bulletin, v. 72, p. 49-72.
J. Pape, and D. G. Roberts, 1985, Depositional sequence mapping as a technique to establish tectonic and stratigraphic framework and evaluate hydrocarbon potential on a passive continental
margin, in O. R. Berg and D. G. Woolverton, eds.. Seismic stratigraphy II: an integrated approach: AAPG Memoir 39, p. 79-91.
Jackson, M. P. A., and W. E. Galloway, 1984, Structural and depositional styles of Gulf Coast Tertiary continental margins: application to hydrocarbon exploration: AAPG Course Notes 25, 226 p.
Kauffman, E. G., 1969, Cretaceous marine cycles of the Western Interior: Mountain Geologist, v. 6, p. 227-245.
Kominz, M. A., 1984, Oceanic ridge volumes and sea-level change, an
error analysis, in J. S. Schlee, ed.. Interregional unconformities and
hydrocarbon accumulation: AAPG Memoir 36, p. 109-127.
Krumbein, W. C , and L. L. Sloss, 1963, Stratigraphy and sedimentation: San Francisco, W. H. Freeman, 660 p.
Leckie, R. M., and P. N. Webb, 1983, Late Oligocene-early Miocene
glacial record of the Ross Sea, Antarctica: evidence from DSDP Site
270: Geology, v. II, p. 578-582.
Loutit, T. S., and J. P. Kennett, 1981, Australian Cenozoic sedimentary
cycles, global sea level changes and deep sea sedimentary record:
Oceanologica Acta Special Volume, p. 45-63.
Meyer, B.L., and M. H. Nederlof, 1984, Identification of source rocks
on wireline logs by density/resistivity and sonic transit time/
resistivity crossplots: AAPG Bulletin, v. 68, p. 121-129.
Miall, A. D., 1984, Principles of sedimentary basin analysis: New York,
Springer-Verlag, 490 p.
1986, Eustatic sea level changes interpreted from seismic stratigraphy: a critique of the methodology with particular reference to
the North Sea Jurassic record: AAPG Bulletin, v. 70, p. 131-137.
Mitchum, R. M., Jr., 1985, Seismic stratigraphic expression of submarine fans, in O. R. Berg and D. G. Woolverton, eds., Seismic stratigraphy II: an integrated approach: AAPG Memoir 39, p. 117-136.
P. R. Vail, and S. Thompson, III, 1977, Seismic stratigraphy
and global changes of sea level, part 2: the depositional sequence as
a basic unit for stratigraphic analysis, in C. E. Payton, ed.. Seismic
stratigraphyapplications to hydrocarbon exploration: AAPG
Memoir 26, p. 53-62.
Molenaar,C. M., 1983, Major depositional cycles and regional correlations of Upper Cretaceous rocks, southern Colorado Plateau and
adjacent areas, in M. W. Reynolds and E. D. Dolly, eds., Mesozoic
paleogeography of west-central United States: SEPM Rocky Mountain Section, p. 201-224.
Morner, N. A., 1980, Eustasy and geoid changes as a function of core/
mantle changes, in N. A. Morner, ed.. Earth rheology, isostasy and
eustasy: New York, John Wiley, p. 535-553.
Mutti, E., 1985, Tlirbidite systems and their relations to depositional
sequences, in G. G. Zuffa, ed., Provenance of arenites: Boston, D.
Reidel,p.65-93.
Newman, W. S., L. F. Marcus, R. Pardi, J. A. Paccione, and S. M.
Tomecek, 1980, Eustasy and deformation of the geoid: 1000-6000
radiocarbon years BP, in N. A. Morner, ed.. Earth rheology, isostasy and eustasy: New York, John Wiley, p. 555-567.
Nummedal, D., and D. J. P. Swift, 1986, The marine transgressive surface as a sequence boundary, a case study of the Upper Coniacian
transgression in the San Juan basin (abs.): Gulf Coast Section
SEPM Research Conference Proceedings, p. 24.
Parkinson, N., and C. Summerhayes, 1985, Synchronous global
sequence boundaries: AAPG Bulletin, v. 69, p. 685-687.
Pitman, W. C , III, 1978, Relationship between eustacy and stratigraphic sequences of passive margins: GSA Bulletin, v. 89, p. 13891403.
and X. Golovchenko, 1983, The effect of sea level change on
the shelf edge and slope of passive margins, in The shelf break: critical interface on continental margins: SEPM Special Publication 33,

142

Genetic Stratigraphic Sequences in Basin Analysis I

p. 41-58.
Poag, C. W., and J. S. Schlee, 1984, Depositional sequences and stratigraphic gaps on submerged United States Atlantic margin, / J. S.
Schlee, ed., Interregional unconformities and hydrocarbon accumulation: AAPG Memoir 36, p. 163-182.
Quinlan, G. M., and C. Beaumont, 1984, Appalachian thrusting, lithospheric flexure, and the Paleozoic stratigraphy of the Eastern Interior of North America: Canadian Journal of Earth Sciences, v. 21,
p. 973-996.
Ru, K., and J. D. Pigott, 1986, Episodic rifting and subsidence in the
South China Sea: AAPG Bulletin, v. 70, p. 1136-1155.
Schumm, S. A., 1977, The fluvial system: New York, John Wiley,
338 p.
Schwan, W., 1980, Geodynamic peaks in alpinotype orogenies and
changes in ocean-floor spreading during Late Jurassic-late Tertiary
time: AAPG Bulletin, v. 64, p. 359-373.
Sloss, L. L., 1963,Sequences in the cratonic interior ofNorth America:
GSABulletin.v. 74, p.93-114.
Stott, D. F., 1984, Cretaceous sequences of the foothills of the Canadian Rocky Mountains: Canadian Society of Petroleum Geologists
Memoir 9, p. 85-107.
Stow, D. A. v., D. G. Howell, and C. H. Nelson, 1985, Sedimentary,
tectonic, and sea-level controls, in A. H. Bouma, W. R. Normark,
and N. E. Barnes, eds.. Submarine fans and related turbidite systems: New York, Springer-Verlag, p. 15-34.
Suter, J. R., and H. L. Berryhill, Jr., 1985, Late Quaternary shelfmargin deltas, northwest Gulf of Mexico: AAPG Bulletin, v. 69, p.
77-91.
Swift, D. J. P., 1968, Coastal erosion and transgressive stratigraphy:
Journal of Geology, v. 76, p. 444-456.
and D. D. Rice, 1984, Sand bodies on muddy shelves: a model
for sedimentation in the Western Interior Cretaceous seaway, North
America, /'n R. W. Tillman and C. T. Siemers, eds., Siliciclastic shelf
sediments: SEPM Special Publication 34, p. 43-62.
Vail, P. R., 1987, Seismic stratigraphy interpretation using sequence
stratigraphy, part 1, seismic stratigraphy interpretation procedure,
in A. W. Bally, ed.. Atlas of seismic stratigraphy, v. 1: AAPG Studies in Geology 27, p. 1-10.
J. Hardenbol, and R. G. Todd, 1984, Jurassic unconformities,
chronostratigraphy, and sea-level changes from seismic stratigra-

phy and biostratigraphy, in J. S. Schlee, ed., Interregional unconformities and hydrocarbon accumulation: AAPG Memoir 36, p.
129-144.
Van Wagoner, J. C , R. M. Mitchum, Jr., H. W. Posamentier, and P. R.
Vail, 1987, Seismic stratigraphy interpretation using sequence stratigraphy, part 2, key definitions of sequence stratigraphy, in A. W.
Bally, ed.. Atlas of seismic stratigraphy, v. I: AAPG Studies in
Geology27, p. 11-14.
Vita-Finzi, C , 1986, Recent earth movements: London, HarcourtBrace, 226 p.
Wanless, H. R., and J. M. Weller, 1932, Correlation and extent of
Pennsylvanian cyclothems: GSA Bulletin, v. 43, p. 1177-1206.
Weimer, R. J., 1970, Rates of deltaic sedimentation and intrabasin
deformation. Upper Cretaceous of Rocky Mountain region, in J. P.
Morgan, ed.. Deltaic sedimentation, modern and ancient: SEPM
Special Publication 15, p. 270-292.
1984, Relation of unconformities, tectonics, and sea-level
changes. Cretaceous of Western Interior, U.S.A., in J. S. Schlee,
ed.. Interregional unconformities and hydrocarbon accumulation:
AAPG Memoir 36, p. 7-35.
Wheeler, H. E., 1960, Early Paleozoic tectono-stratigraphic patterns in
the United States: 21st International Geological Congress Reports,
part 8, p. 47-56.
Wiltschko, D. v., and J. A. Dorr, Jr., 1983, Timing of deformation in
overthrust beh and foreland of Idaho, Wyoming, and Utah: AAPG
Bulletin, V. 67, p. 1304-1322.
Winker, C. D., 1979, Late Pleistocene fluvial-deltaic deposition, Texas
coastal plain and shelf: Master's thesis. University of Texas at Austin, Austin, Texas, 187 p.
1982, Cenozoic shelf margins, northwestern Gulf of Mexico
basin: Gulf Coast Association of Geological Societies Transactions,
V. 32, p. 427-448.
1984, Clastic shelf margins of the post-Comanchean Gulf of
Mexico: implications for deep water sedimentation: Gulf Coast Section SEPM 5th Annual Research Conference, p. 109-117.
Zeller, E. J., 1964, Cycles and psychology: Kansas Geological Survey
Bulletin 169, p. 631-636.
Zoback, M. L., and M. Zoback, 1980, State of stress in the conterminous United States: Journal of Geophysical Research, v. 85, p.
6113-6156.

You might also like