You are on page 1of 275

AAPG Studies in Geology #27

VOLUME 2

ATLAS OF SEISMIC
STRATIGRAPHY
THE BLIND MEN AND THE ELEPHANT
It was six men of Indostan
To learning much inclined,
Who went to see the Elephant
(Though all of them were blind),
That each by observation
Might satisfy his mind.

The Fifth who chanced to touch the ear


Said: E en the blindest man
Can tell what this resembles most;
Deny the fact who can,
This marvel of an Elephant
Is very like a fan!

The First approached the Elephant,


And happening to fall
Against his broad and sturdy side,
At once began to bawl:
God bless me! but the Elephant
Is very like a wall!

The Sixth no sooner had begun


About the beast to grope,
Than, seizing on the swinging tail
That fell within his scope,
I see, quoth he, the Elephant
Is very like a rope!

The Second, feeling of the tusk,


Cried, Ho! what have we here
So very round and smooth and sharp?
To me tis mighty clear
This wonder of an Elephant
Is very like a spear!

And so these men of Indostan


Disputed loud and long,
Each in his own opinion
Exceeding stiff and strong,
Though each was partly in the right
And all were in the wrong!

The Third approached the animal,


And happening to take
The squirming trunk within his hands,
Thus boldly up and spake:
I see, quoth he, the Elephant
Is very like a snake!

MORAL
So oft in theologic wars,
The disputants, I ween,
Rail on in utter ignorance
Of what each other mean,
And prate about an Elephant
Not one of them has seen!

The Fourth reached out an eager hand,


And felt about the knee.
What most this wondrous beast is like
Is mighty plain, quoth he;
Tis clear enough the Elephant
Is very like a tree !

John Godfrey Saxe


(1816-1887)

The American Association of Petroleum Geologists

AAPG STUDIES IN GEOLOGY #27


VOLUME 2

ATLAS OF
SEISMIC STRATIGRAPHY
EDITED BY
A.W. BALLY
Rice University, Houston, Texas

Published by The American Association of Petroleum Geologists,Tulsa, Oklahoma 74101, U.S.A.

ACKNOWLEDGMENT

PUBLISHERS NOTE

We thank the following organizations for their generous contributions to


the content of this atlas:
AGIP (Societa per Azioni-PIGE), Milan, Italy
Amoco Production Company, Houston, Texas and New Orleans, Louisiana
ARCO Oil and Gas Company, Plano, Texas
BHP Petroleum Pty Ltd., Melbourne, Australia
Bohai Oil Company, Tianjin, Peoples Republic of China
Chevron U.S.A. Inc., San Ramon, California
Dalhousie University, Department of Oceanography, Halifax, Nova Scotia
The Energists, Houston, Texas
Exxon Production Research Company, Houston, Texas
Geoquest International Inc., Houston, Texas
Gulf Research and Development Company, Houston, Texas
Institut Francais du Petrole (IFP), Paris, France
Koninklijke/Shell Exploratorie en Produktie Laboratorium, Shell Research
B.V., The Hague, Netherlands
Lamont-Doherty Geological Observatory of Columbia University,
Palisades, New York
Charles Leung, Consultant, Houston, Texas
Allen Lowrie, Consultant, Picayune, Mississippi
L.D. Meckel and Company, Houston, Texas
Marathon Oil Company, Littleton, Colorado
NOPEC Geo Services, A.S, Oslo, Norway
Petrobras, Rio De Janeiro, Brazil
Petro-Canada Resources, Calgary, Alberta, Canada
Sarawak Shell Berhad, Sarawak, Malaysia
Schlumberger-Doll Research, Ridgefield, Connecticut
Scripps Institution of Oceanography, La Jolla, California
Shell Oil Company, Houston, Texas
Texas A & M University, College Station, Texas
Total-CFP (Compagnie Francais des Petroles), Paris, France
Universite, Bordeaux I, Institut de Geologie du Bassin dAquitaine, Talence,
France
University of Ferrara, Instituto di Geologia, Ferrara, Italy
University of Houston, Houston, Texas
University of Miami, Miami Beach, Florida
University of Texas at Austin, Institute for Geophysics, Austin, Texas
U.S. Geological Survey, Woods Hole, Massachusetts
Western Research, Division of Western Geophysical, Houston, Texas

The American Association of Petroleum Geologists gratefully


acknowledges the hundreds of geologists and geophysicists who gave their
time and interpretations to make this comparative atlas possible, and the
scores of companies and organizations who allowed their personnel to
participate. We list contributing organizations to the left, under
Acknowledgment.
The large-size format is designed to be big enough to reproduce each
seismic line at a useful size, yet the book is small enough (like the atlas of
structural styles, which preceded this volume set) to be conveniently stored
in a standard-size filing cabinet drawer.
The overwhelming response by seismic line contributors has given us a
wealth of examples in this collection. We especially are grateful to workers
at Shell and Exxon for key parts of the first volume, dealing with
interpretation and the self-teaching of interpretation.
To accommodate the number of pages involved in such a compilation,
we broke the work into three volumes. Volume One discusses interpretation
and interpretation procedure, geometric and facies analysis, and basin-fill
analysis, then closes with method-oriented papers, which discuss
interpretation through case history. Volume Two carries the case history
style to Rift Systems (three examples), and Passive Margins (25 examples)
from around the world. Volume Three offers case histories from the Gulf of
Mexico (13 examples), The Foredeep (7 examples), Active Margins (8
examples), and Deep Sea Stratigraphy (2 examples).
Lastly, we thank A.W. Bally for his foresight in perceiving the need for
such a volume set, and for his work in compilation. We hope weve
developed a useful teaching tool for classroom use or for personal study.
The Tables of Contents for all three volumes are presented here so that
the reader can consider the subsequent treatment given these terranes.
AAPG Publications
Tulsa, Oklahoma

iv

FOREWORD
I am confident that someday the concept of geological time will be
acclaimed as one of the more wonderful contributions from natural science.
With these words, Claud Albritton (1980) begins his most readable tale
of the Abyss of Time. Albritton tells us that Nicholas Steno in 1669
enunciated the principles of superposition and stratigraphic continuity, and
Albritton offers a reproduction of Stenos first illustration attempting to
explain the separation of two sequences by an unconformity. Reading
further in Albrittons book, we are told how William Smith, in 1799,
parlayed the principle of faunal succession into a stratigraphic table and
finished producing a map of England that showed the distribution of various
strata. Stratigraphy thus became a foundation of modern geology. Today,
seismic stratigraphy has brought about a revival of the thought processes
that were initiated by Nicholas Steno and William Smith.
It is not my purpose here to review further the history of stratigraphy.
Others have provided fine summaries (e.g., Dott, 1971; and Conkin and
Conkin, 1984). In recent decades, stratigraphy has branched out in
numerous specialized fields, such as the various fields of sedimentology,
quantitative stratigraphic methodology, and paleontological stratigraphies,
to name a few. However, traditional stratigraphy, with some justification,
has become more and more concerned with bringing order into matters of
stratigraphic codification, correlations and the formal blessing of type sections (see
Hedberg, 1976). This drift toward specialization and codification appeared for a time
to diminish the interest of many research scientists in general stratigraphy. There were
notable exceptions, however; like L. Sloss (1963, 1972) and H.E. Wheeler (1963),
who founded modern sequence stratigraphy and proposed continenteven
worldwidesequence correlations.
Fisher and Arthur (1977) and Fisher (1981, 1982, and 1983) devoted much effort
to pinning down and explaining stratigraphic cycles and climatic oscillations. A
number of symposia on cyclic sedimentation have been published (Merriam, 1964;
Duff et al., 1967, Schwarzacher, 1975; and Einsele and Seilacher, 1982).
A major breakthrough revitalized stratigraphy when Vail et al.
(1977), following the earlier path of Sloss and Wheeler, opened up
entirely new vistas of global stratigraphy by introducing for the
first time a rigorous stratigraphic interpretation of seismic
reflection profiles. Vail and his colleagues at Exxon proposed
global sea-level changes that could be mapped by careful seismic
stratigraphic analysis. As would be expected, there was and still is a
significant debate questioning the sea-level interpretation of Vails
curve. I will not enter into that debate, because it is the purpose of
this atlas to make material available that will permit teachers and
students to come up with their own interpretations and to make up their own
minds.

Perhaps the most fundamental assumption of the seismic stratigraphic


method is that a seismic reflector in most cases, and for all practical
purposes, is a time line. This, of course, can be questioned on theoretical
grounds. After all, reflections are due to impedance contrasts between
layered rocks, and we know of reflectors that correspond to diagenetic
fronts or to fluid contacts; we are also familiar with hydrate reflectors that
often mimic the sea floor. Despite these and other obvious exceptions, it is
by now clear that in sedimentary, layered rocks, seismic reflectors are
indeed very fine approximations to time lines. More often than not,
reflection profiles offer finer stratigraphic subdivisions than can be obtained
by paleontological studies. Obviously, we need all: reflection profiles, well
logs, outcrops sections and paleontological calibration; but generally
speaking, if the number of reflectors exceeds the number of paleozones we
have more detailed time-stratigraphic resolution. More importantly, perhaps,
reflection profiles allow us to follow strata packages over long distances and
wide subsurface areas, thus permitting regional and supraregional
correlations that often are impossible to obtain by conventional field work
and/or subsurface stratigraphy.
Why publish a seismic stratigraphic atlas? If we are going to continue
our debate on seismic stratigraphy, we need to involve more earth scientists.
This means that seismic profiles showing beautiful stratigraphy should be
accessible to students and teachers. It is important that each profile have an
uninterpreted version together with an interpretation. Interpretations involve
judgments, and judgments generate debate; debate in turn generates new
hypotheses to be tested. Clearly, that whole process simply has to start with
the availability of pristine data. Learned journals today tend to show
miniaturized seismic profiles that are covered up with interpretations,
preventing the reader from separating fact, factoids and fiction. In this atlas
we attempt to minimize that problem by including many uninterpreted
profiles.
We took advantage of the momentum we gained with the
publication of the three-volume Seismic Expression of Structural
Styles (Bally, 1983). These volumes turned out to be rather
popular. It was particularly gratifying to see similar projects spring
up simultaneously elsewhere, such as the Seismic Atlas of
Australian and New Zealand Sedimentary Basins (Skilbeck and
Lennox, 1984) and the RMAG-DGS volume Seismic Exploration
of the Rocky Mountain Region (Gries and Dyer, 1985). Both of
these beautiful compilations, as well as the earlier AAPG Structural
Styles volumes, contain numerous other examples that could be
used to further illustrate seismic stratigraphy.

ORGANIZATION OF THE ATLAS


There is no easy and slick way to organize the fine contributions
offered by many authors for this atlas. No lofty classification principles are
available to structure such varied material. I therefore simply tried to
arrange the varying subjects into related clusters.
Two very significant contributions constitute the introductory chapters: I
am elated to start this series with an update on seismic stratigraphic
principles, prepared by P. Vail and colleagues. The reader should note that
these have evolved considerably since the publication of AAPG Memoir 26
(Payton, 1977). The second part of the introduction contains an atlas of
seismic stratigraphy, prepared by Shell Oil Company. We felt that the
outstanding graphics of that atlas were particularly important for teachers
and students.
The next batch of chapters includes a diverse group of method-oriented
papers. Of course, there is much more material available on seismic
stratigraphic methodology. We refer the reader to AAPG Memoirs 26
(Payton, 1877), 36 (Schlee, 1984), and 39 (Berg and Woolverton, 1985). In
the same context, the geologically oriented reader is referred to the
following books on reflection seismology: Anstey, 1982; McDonald, et al.,
1983; Fitch, 1976; Neidell, 1980; Sheriff, 1973; Sheriff and Geldart, 1982,
1983; Tucker, 1982; Tucker and Yorston, 1973; and Waters, 1981.
The remainder and main part of the atlas is grouped into several major
topics: rift systems, passive margins, Gulf of Mexico, foredeeps, active
margins, and deep sea stratigraphy. We resisted the temptation to group the
articles into clastic papers and carbonate papers, because both clastics and
carbonates illustrate the fundamentals of seismic stratigraphyalbeit often
with different expressions.
I point to the carbonate papers because, in very recent years, seismic
stratigraphy has begun to greatly influence carbonate sedimentology. Earlier
carbonate specialists proposed facies models and sequences by comparing
recent with ancient carbonates, by juxtaposing various facies in models, and
by studying diagenetic phenomena. In the process of this important work,
carbonate stratigraphy was perhaps neglected, with the net effect that facies
that were not synchronous were erroneously juxtaposed. Some years ago, at
a memorable Penrose conference organized by Jim Wilson and Bruno
dArgenio on the Island of Capri, we were reminded of the old tale of the
blind men who tried to recognize an elephant but were unable to do so
because each had felt only parts of the animal. At that time, I reminded the
participants that one could actually see the whole elephant on seismic
profiles. Reacting to my rhetorical hyperbole, Lloyd Pray sent me a Xerox
of an elephant pasted on a seismic profile, which inspired our present atlas
cover. It would be comforting to know that, in keeping with J.J. Audubons
famous Elephant Folio of the Birds of North America, our atlas volumes
would in the future be referred to as the Elephant Folio of Stratigraphy.

What Next?
As with the structural atlas, we again received numerous suggestions
that the material presented in the seismic stratigraphic atlas be accompanied
by a text on modern stratigraphy. P. Vail is planning to prepare such a text,
but obviously to do this right will take time. There also have been
suggestions to prepare sets of films for classroom exercises in seismic
stratigraphy. Here again, P. Vail and I are beginning to assemble the best
material for an exercise set, but we will need more time to bring that project
to a conclusion.

REFERENCES
Albritton, C.C., 1980, The abyss of time: Freeman, San Francisco, 251 p.
Anstey, N.A., 1982, Simple seismics: Boston, IHRDC, 168 p.
Bally, A. W., ed., 1983 and 1984, Seismic expression of structural stylesa picture
and work atlas: American Association of Petroleum Geologists, Studies
#15, 3 volumes.
Brown, A.R., 1986, Interpretation of three-dimensional seismic data: American
Association of Petroleum Geologists, Memoir 42, 194 p.
Berg, O.R., and D.G. Woolverton, eds., 1985, Seismic stratigraphy IIan
integrated approach: American Association of Petroleum Geologists,
Memoir 39, 276 p.
Conkin, B.M., and J.E. Conkin, 1984, Stratigraphy, foundations and concepts: New
York, Van Nostrand, Benchmark series, 365 p.
Dott, R.H., and R.L. Batten, 1971, Evolution of the earth: McGraw-Hill, 649 p.
Duff, P., McL. D., A. Hallam, and E.K. Walton, eds., 1967, Cyclic sedimentation,
developments in sedimentation, developments in sedimentology:
Amsterdam, Elsevier, 280 p.
Einsele, G., and A. Seilacher, eds., 1982, Cyclic and event stratification: Berlin,
Springer Verlag, 536 p.
Fisher, A.G., 1981, Climatic oscillations in the biosphere, in Crisis in ecological
and evolutionary time, M. Niteckie, ed.: New York, Academic Press, p.
103-131.
Fisher, A. G., 1982, Long term climatic oscillations recorded in stratigraphy, in
Climate in earth history, W.H. Berger and J.C. Crowell, eds.: Washington,
D.C., Natural Academy of Sciences Press, p. 97-104.
Fisher, A. G., 1983, the two Phanerozoic supercycles, in Catastrophes in earth
history: the new uniformitarianism, W. Berggren and J. van Couvering,
eds.: Princeton, Princeton University Press, p. 129-150.
Fisher, A. G., and M. Arthur, 1977, Secular variations in the pelagic realm, in Deep
watercarbonate environments, H.E. Cook and P. Enos, eds.: Society of Economic
Paleontologists and Mineralogists Special Publication 25, p. 19-50.
Fitch, A.A., 1976, Seismic reflection interpretation (United Nations Development
Programme, Quito, Ecuador): Berlin, Gebrueder Borntraeger, XXI, 148 p.
Gries, R.R., and R.C. Dyer, 1985, Seismic exploration of the Rocky Mountain region:
Rocky Mountain Association Geologists and Denver Geophysical Society,

299 p.
Hedberg, H.D., ed., 1976, International stratigraphic guide-a guide to

stratigraphic classification, terminology, and procedure: New York, John


Wiley, 200 p.
Holland, H.D., and A.F. Trendall, eds., 1984, Patterns of change in earth evolution:
Dahlem Konferenzen: Berlin, Springer Verlag, 431 p.
McDonald, J.A., G.H.F. Gardner, and F. J. Hiltermann, eds., 1983, Seismic studies
in physical modeling: Boston, IHRDC, 256 p.
Merriam, D.F., ed., 1964, Symposium on cyclic sedimentation: Kansas Geological
Survey Bulletin 169, 636 p., vol. 1 and 2.
Neidell, N.S., 1980, Stratigraphic modeling and interpretation: geophysical
principles and techniques: American Association of Petroleum Geologists
Continuing Education Course Notes Series 13, 145 p.
Payton, C.E., ed., 1977, Seismic stratigraphyapplications to hydrocarbon
exploration:American Association of Petroleum Geologists Memoir 26, 516 p.
Schlee, J.S., ed., 1984, Interregional unconformities and hydrocarbon
accumulation: American Association of Petroleum Geologists Memoir 36,
184 p.
Schwarzacher, W., 1975, Sedimentation models and quantitative stratigraphy:
Developments in Sedimentology 19: Amsterdam, Elsevier, 383 p.
Sheriff, R.E., 1973, Encyclopedic dictionary of exploration geophysics: Tulsa,
Society of Exploration Geophysicists, 276 p.
Sheriff, R.E., 1980, Seismic stratigraphy: Boston, IHRDC, 227 p.
Sheriff, R.E., and L. P. Geldart, 1982 and 1983, Exploration seismology (2 vols.),
volume 1: history, theory and data acquisition, 1982, 243 p.; volume 2:
data-processing and interpretation, 221 p: Cambridge University Press.
Skilbeck, C.G., and M.L. Lennox, eds., 1984, the seismic atlas of Australian and
New Zealand sedimentary basins: Syndey, The Earth Resources
Foundation, University of Sydney, 301 p.
Sloss, L.L., 1963, Sequences in the cratonic interior of North America: Geological
Society America Bulletin, v. 74, p. 93-113.
Sloss, L.L., 1972, Synchrony of Phanerozoic sedimentary-tectonic events of the
North American craton and the Russian platform: International Geological
Congress, 24th Session Section 4, p. 24-32.
Sloss, L.L., 1976, Areas and volumes of cratonic sediments, western North
America and Eastern Europe: Geology, v. 4, p. 272-276.
Tucker, P. M., 1982, Pitfalls revisited: Tulsa, Society of Exploration Geophysicists,
23 p.
Tucker, P. M., and H.Y. Yorston, 1973, Pitfalls in seismic interpretation: Tulsa,
Society of Exploration Geophysicists Monograph Series No. 2, 56 p.
Vail, P.R., et al., 1977, Seismic stratigraphy and global changes of sea level (a
series of papers), in C.E. Payton, ed., Seismic stratigraphyapplication to
hydrocarbon exploration: American Association of Petroleum Geologists
Memoir 26, p. 49-212.
Waters, K.H., 1981, Reflection seismology, 2nd ed.: New York, John Wiley, 465 p.
Wheeler, H.E., 1963, Post-Sauk and pre-Absaroka Paleozoic stratigraphic patterns
in North America: American Association of Petroleum Geologists Bulletin,
v. 47, p. 1497-1526.

vi

Table of ContentsVolume 2
Acknowledgement..................................................................................................... iv
Foreword ................................................................................................................... v
List of Figures and Tables ......................................................................................... viii
Swift Systems
Synrift and postrift sequences in the northern North Sea .........................................
T. Pettersen, P.T. Svela and H. Carstens

Structural and stratigraphic framework of the Jeanne DArc basin, Grand Banks ... 14
H.J. Welsink and A. Tankard
The seismic reflection characteristic and oil-gas-bearing condition of main sand
sedimentary bodies in Eogene faultingLake Basin in Bohai Bay ......................... 22
Jiang Xijiang
Passive Margins
Interpretation of West Africa, line C* ....................................................................... 37
R. Sheriff
Listric faults, offshore Morocco................................................................................ 45
D.B. Macurda, Jr.
An update of U.S. Geological Survey seismic reflection line 25 across the
New Jersey shelf, slope and upper rise* ................................................................... 48
J.S. Schlee, C.W. Poag and J.A. Grow
Baltimore Canyon trough, Mid-Atlantic OCS: seismic stratigraphy of
Shell/Amoco/Sun wells*........................................................................................... 51
R.N. Erlich, K.P. Maher, G.A. Hummel, D.G. Benson,
G.J. Kastritis, H.D. Linder, R.S. Hoar and D.H. Neeley
Tertiary depositional sequences, offshore New Jersey and Alabama........................ 67
S.M. Greenlee

USGS line 19 across the Georges Bank basin* ........................................................ 81


K.D. Klitgord, J.S. Schlee and J.A. Grow

Para-Maranhao basinBrazil*................................................................................. 179


Petrobras Exploration Department

Contourites and volcanics, Georges Bank, New England......................................... 84


D.B. Macurda, Jr.

Potiguar basin*.......................................................................................................... 185


Petrobras Exploration Department

Shelf-to-basin correlations off eastern Canada: developing s seismic


stratigraphic framework in the northern Newfoundland basin.................................. 88
K.J. Meador, J.A. Austin, Jr. and D.F. Dean

Jequitinhonha basin*................................................................................................. 191


Petrobras Exploration Department

Aggrading and prograding infill of buried Cenozoic seaways,


northwestern Great Bahama Bank ............................................................................ 97
G. Eberli and R. Ginsburg
Deep clastic carbonate deposits of the Bahamascomparison with
Mesozoic outcrops of the Vercors and of the Vocontian trough*.............................. 104
C. Ravenne, P. LeQuellec, P. Valery and R. Vially
Eroded carbonate platform marginthe Blake esc arpment off
southeastern United States* ...................................................................................... 140
W.P. Dillon, A.M. Trehu and P.C. Valentine

Outcrop models for seismic stratigraphy: examples from the Triassic of the
Dolomites* ................................................................................................................ 194
A. Bosellini
Regional seismic stratigraphic analysis of Upper JurassicLower Cretaceous
carbonate depositional sequences, Neuquen basin, Argentina.................................. 206
R.M. Mitchum, Jr. and M.A. Uliana
Tertiary hiatuses in Western Approaches .................................................................. 213
C. Ravenne, C. Muller and L. Montadert

Erosion of the southern Florida escarpment* ........................................................... 149


W. Corso, R.T. Buffler and J.A. Austin, Jr.

Seismic stratigraphic features of the Porcupine basin, offshore Ireland ...................240


D.B. Macurda, Jr.
Mass slides and turbidite type deposits recognized by offshore seismic prospecting: Cap Ferret depression and at the outcrop: Gres dAnnot series ................ 248
C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Seismic stratigraphy of carbonate platform sediments, southwest Florida*


D.B. Macurda, Jr.

Submarine canyons, contourites, and volcanics, Mozambique................................. 265


D.B. Macurda, Jr.

Seismic stratigraphy of the Exmouth plateau ........................................................... 163


R.D. Erskine and P.R. Vail

Relative sea level changes and depositional modes of the shelf and deep sea
fan of the Indus ......................................................................................................... 270
C. Ravenne, F. Courmes and J.P. Esteve

Cenozoic carbonate banks, Foz Do Amazonas basin, northeastern Brazil* ............. 174
M.J.N. Brouwer and M.M. Schwander

* Indicates papers that focus on carbonate stratigraphy

vii

AAPG Atlas of Seismic Stratigraphy


List of Figures and Tables
Figure 1. Index map showing four study areas described in this chapter. ......................................................................................................................................................................................................................................................................................................................................................................................
Figure 2. Index map showing location of line NOD 4-84-30, Tampen Spur, and relation to bounding features, northern North Sea. .........................................................................................................................................................................................................................................................................................
Figure 3. Index map showing location of lines NOD 4-84-19, Sogn graben east, and NOD 3-84-39, Horda Platform, and their relation to bounding features, northern North Sea................................................................................................................................................................................................................
Figure 4. Index map showing location of line CNST 82-06, Egersund subbasin, and its relation to bounding features, northern North Sea. .............................................................................................................................................................................................................................................................................
Figure 5. Uninterpreted migrated section of line NOD 4-84-30 across Tampen Spur (see Figure 2 for location), northern North Sea. Figure 6 shows authors interpretation.........................................................................................................................................................................................................................
Figure 6. Interpreted migrated section of line NOD 4-84-30 (see Figure 5)..................................................................................................................................................................................................................................................................................................................................................................................
Figure 7. Uninterpreted migrated section of line NOD 4-84-19 across Sogn graben east (see Figure 3 for location), northern North Sea. Figure 8 shows authors interpretation...................................................................................................................................................................................................................
Figure 8. Interpreted migrated section of line NOD-4-84-19 (see Figure 7). ................................................................................................................................................................................................................................................................................................................................................................................
Figure 9. Uninterpreted migrated section of line NOD 3-84-39 across Horda Platform (see Figure 3 for location), northern North Sea. Figure 10 shows authors interpretation. ..................................................................................................................................................................................................................
Figure 10. Interpreted migrated section of line NOD 3-84-39 (see Figure 9)................................................................................................................................................................................................................................................................................................................................................................................
Figure 11. Uninterpreted migrated section of line CNST 82-06 across Egersund subbasin (see Figure 4 for location), northern North Sea. Figure 12 shows authors interpretation..............................................................................................................................................................................................................
Figure 12. Interpreted migrated section of line CNST 82-06 (see Figure 11). ..............................................................................................................................................................................................................................................................................................................................................................................
Figure 1. The Grand Banks are the continental shelf offshore from Newfoundland. It is separated from the Scotian and Labrador shelves by the Newfoundland and Charlie Gibbs fracture zones, respectively. Internally, major strike-slip faults divide the Grand Banks into southern (S), central (C), and
northern (N) provinces. On a smaller scale, transfer faults accommodate different amounts and rates of extension. Basin geometries are based on seismic and gravity data. The ocean-continental boundary (OCB) is after Masson and Miles (1984), the J-anomaly is after Sullivan and Keen (1978). ..................
Figure 2. Tectono-stratigraphic column for the Jeanne dArc basin, showing major episodes in basin evolution separated by prominent unconformities. These unconformity-bounded sequences match the megacycles of rift and postrift subsidence as well as various phases of ocean opening about the
Grand Banks: separation of Africa and Aalemian transgression (Whale limestone); Blake Spur Magnetic Anomaly and spreading center shift (late Callovian extension); separation of Grand Banks from Iberia (Aptian unconformity); separation of Flemish Cap from Galicia margin (pre-Cenomanian
unconformity); separation of the European-Greenland plate (basal-Tertiary unconformity). .......................................................................................................................................................................................................................................................................................................................................................
Figure 3. Different structural styles have combined to dissect the Jeanne dArc basin into a mosaic of small fault-bounded blocks. Two major fault trends are dominant: southwest-to-northeast basin bounding faults; southeast-to-northwest cross-basin faults, the major ones (thick lines) being transfer
faults. A third trend exists on a smaller scale: conjugate Riedel shears obliquely connecting transfer faults; potential tear faults trending parallel or slightly obliquely to basin-bounding faults..........................................................................................................................................................................................
Figure 4. Geological depth section of Profile A (Figures 5 and 6) showing major unconformity-bounded sequences and lithological units. Lithologies come from various wells projected on Profile A. ...........................................................................................................................................................................
Figure 5. Migrated seismic section of Profile A (see Figure 3 for location), Jeanne dArc basin, Grand Banks. ........................................................................................................................................................................................................................................................................................................................
Figure 6. Interpretation of Profile A (Figure 5). .............................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 7. Migrated seismic section of Profile B (see Figure 3 for location), Jeanne DArc basin, Grand Banks..........................................................................................................................................................................................................................................................................................................................
Figure 8. Interpretation of Profile B (Figure 7). .............................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 1. Index map of major structural features offshore, Bohai Bay, PRC, and location of seismic lines; contour in km on the base of the Tertiary. .............................................................................................................................................................................................................................................................
Figure 2. Paleogene lacustrine sand units in Bohai basin, northeast China (uninterpreted). Line 1169.5, BZ-28 field area.........................................................................................................................................................................................................................................................................................................
Figure 3. Paleogene lacustrine sand units in Bohai basin, northeast China (interpreted). Line 1169.5, BZ-28 field area.............................................................................................................................................................................................................................................................................................................
Figure 4. Paleogene lacustrine sand units in Bohai basin, northeast China (uninterpreted). Line 1077, HZ-28 number 2 fan. ....................................................................................................................................................................................................................................................................................................
Figure 5. Paleogene lacustrine sand units in Bohai basin, northeast China (interpreted). Line 1077, HZ number 2 fan. .............................................................................................................................................................................................................................................................................................................
Figure 6. Paleogene lacustrine sand units in Bohai basin, northeast China (uninterpreted). Line 67, Paleo-Luanhe fan delta. ....................................................................................................................................................................................................................................................................................................
Figure 7. Paleogene lacustrine sand units in Bohai basin, northeast China (interpreted). Line 67, Paleo-Luanhe fan delta. ........................................................................................................................................................................................................................................................................................................
Figure 8. Paleogene lacustrine sand units in Bohai basin, northeast China (uninterpreted). W-E line 134, LZW Sag. ................................................................................................................................................................................................................................................................................................................

3
3
4
5
6
7
8
9
10
11
12
13
14
15
15
16
18
19
20
21
23
24
25
26
27
28
29
30

viii

Figure 9. Paleogene lacustrine sand units in Bohai basin, northeast China (interpreted). W-E line 134, LZW Sag. ....................................................................................................................................................................................................................................................................................................................
Figure 10. Paleogene lacustrine sand units in Bohai basin, northeast China (uninterpreted). SW-NE line 138, deep lake progradation. ....................................................................................................................................................................................................................................................................................
Figure 11. Paleogene lacustrine sand units in Bohai basin, northeast China (interpreted). SW-NE line 138, deep lake progradation. ........................................................................................................................................................................................................................................................................................
Figure 1. Interpretation of West Africa Line C (shown in AAPG Memoir 26, p. 157). .................................................................................................................................................................................................................................................................................................................................................................
Figure 2. Original section used in AAPG Seismic Stratigraphy Schools.......................................................................................................................................................................................................................................................................................................................................................................................
Figure 3. Section (from Figure 2) reprocessed by Geophysical Development Corporation. .........................................................................................................................................................................................................................................................................................................................................................
Figure 4. Interpretation of the reprocessed section using the same age identification as in Figure 1. ...........................................................................................................................................................................................................................................................................................................................................
Figure 5. Part of the original section, showing top-lap indications that are easier to see than on the reprocessed section. Events immediately above the top-laps are deleted for clarity. .......................................................................................................................................................................................................
Figure 6. Reprocessed section showing locations of velocity analyses and contours of equal stacking velocity. The velocity analyses shown in Figures 7, 8, and 9 are indicated along the top............................................................................................................................................................................................
Figure 7. Interval velocities calculated from the velocity analyses using the Dix equation for the three surveys over the Jurassic shelf. The solid curve shows the expected velocities for a Tertiary clastic section subjected only to the pressure of the overburden (from Sheriff and Geldart, 1983, p. 8).
See Figure 6 for location. ...............................................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 8. Interval velocities for the three surveys over the intermediate shelf. See Figure 6 for locations. ..................................................................................................................................................................................................................................................................................................................................
Figure 9. Interval velocities for six of the velocity surveys over the left end of the line where the water is deepening. A. Plotted with respect to a sea-level datum; B. plotted with respect to a sea-floor datum. ................................................................................................................................................................
Figure 10. Interpretation of flow features at the left end of the line...............................................................................................................................................................................................................................................................................................................................................................................................
Figure 11. Display of the reprocessed section linear in depth rather than linear in time, and with an aspect ratio (vertical to horizontal scale) of 2:1...............................................................................................................................................................................................................................................................
Figure 1. Tarfaya basin, offshore Morocco.....................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 2. Tarfaya basin, offshore Morocco, southwest of Figure 1................................................................................................................................................................................................................................................................................................................................................................................................

42
42
42
43
44
46
47

Table 1. Inferred and documented (through drilling) reflector ages and stratigraphic relations. ...................................................................................................................................................................................................................................................................................................................................................

48

Figure 1. Annotated depth section across the northern Baltimore Canyon Trough, seaward of New Jersey. Major dislocation of reflections near shotpoint 2900 is caused by line cessation while the ship circled to avoid an obstacle. Nearby wells are projected to the line at the points and
to the depths indicated. ...................................................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 1. Index map to study area: Lower Cretaceous-Jurassic reef trend, eastern North America.
Figure 2. Locations of wells in the Baltimore Canyon Trough; index to seismic lines..................................................................................................................................................................................................................................................................................................................................................................
Figure 3. Stratigraphic column, synthetic seismic strip, and logs from Shell No. 1 well, Block 587 (Civet). Albian-upper Necomian section is undifferentiated in this well..........................................................................................................................................................................................................................
Figure 4. Seismic line A-A; uninterpreted migrated time. ............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 5. Seismic line A-A; interpreted migrated time. ................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 6. Seismic line A-A; uninterpreted migrated depth............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 7. Seismic line A-A; interpreted migrated depth................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 8. Enlargement of seismic line A-A; interpreted migrated line..........................................................................................................................................................................................................................................................................................................................................................................................
Figure 9. Seismic line B-B; uninterpreted migrated time. ............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 10. Seismic line B-B; interpreted migrated time. ..............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 11. Seismic line B-B; uninterpreted migrated depth. ........................................................................................................................................................................................................................................................................................................................................................................................................
Figure 12. Seismic line B-B; interpreted migrated depth. ............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 13. Schematic diagram showing generalized depositional environments and seismic horizons. Note that structure was removed for clarity. ................................................................................................................................................................................................................................................................
Figure 14. Development models for Kimmeridgian-lower Neocomian shelf margin. Adapted from McIlreath and James (1979)..............................................................................................................................................................................................................................................................................................
Figure 15. Facies model for Tithonian-lower Neocomian carbonate platform. Adapted from Turnsek et al. (1981). ...................................................................................................................................................................................................................................................................................................................
Figure 16. Generalized coastal onlap curves for the Civet and Rhino wells compared to the curve of Vail et al. (1984). ............................................................................................................................................................................................................................................................................................................
Figure 1. Location of seismic sections illustrated, offshore wells, and positions of Neogene depositional shelf edges, offshore New Jersey. ............................................................................................................................................................................................................................................................................
Figure 2. Location of seismic sections illustrated, offshore wells used in study, and positions of Lower Cretaceous and Neogene shelf edges, offshore Alabama...........................................................................................................................................................................................................................................

31
32
33
37
38
39
40
41
41

49
51
53
54
55
56
57
59
60
61
62
63
64
65
65
65
67
68

ix

Figure 3. Geohistory diagrams of (a) the COST B-2 well (offshore New Jersey) and (b) the Shell Main Pass 154 well (offshore Alabama). The lower curve represents total subsidence of the basement through time; the upper curve represents total subsidence corrected for sediment load and
compaction effects. The uppermost curve represents a long-term sea level curve (Haq et al., 1987) used as a datum, and the shaded area represents paleowater depth interpretation............................................................................................................................................................................................................
Figure 4. Stratigraphic chart including a Tertiary coastal onlap chart from Haq et al. (1987) and coastal onlap from offshore New Jersey and offshore Alabama study areas. .......................................................................................................................................................................................................................
Figure 4. (continued). .....................................................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 5A. Uninterpreted seismic section and interpreted seismic section, offshore New Jersey. ................................................................................................................................................................................................................................................................................................................................................
Figure 5B. Chronostratigraphic diagram of line shown in 5A, offshore New Jersey.....................................................................................................................................................................................................................................................................................................................................................................
Figure 6A. Uninterpreted seismic section and interpreted seismic section, passing through the COST B-2 well location and Exxon 684-1 well, offshore New Jersey....................................................................................................................................................................................................................................
Figure 6B. Miocene chronostratigraphic diagram of the sections shown in Figure 6A.................................................................................................................................................................................................................................................................................................................................................................
Figure 7A. Uninterpreted seismic section and interpreted seismic section, offshore Alabama. ....................................................................................................................................................................................................................................................................................................................................................
Figure 7B. Chronostratigraphic diagram of section shown in Figure 7A. .....................................................................................................................................................................................................................................................................................................................................................................................
Figure 8A. Uninterpreted seismic section, offshore Alabama........................................................................................................................................................................................................................................................................................................................................................................................................
Figure 8B. Interpreted seismic section, offshore Alabama.............................................................................................................................................................................................................................................................................................................................................................................................................
Figure 8C. Chronostratigraphic diagram of the section shown in Figures 8A and 8B...................................................................................................................................................................................................................................................................................................................................................................

69
70
71
72
73
74
75
76
77
78
79
80

Table 1. Inferred reflector ages.......................................................................................................................................................................................................................................................................................................................................................................................................................................................

82

Figure 1. Tectonic elements for the Georges Bank-Gulf of Maine region, including faults and graben structures on the platforms and along the landward edge of the Georges Bank basin, location of line 19, salt-diapir province along the seaward edge of Georges Bank and fracture zones (FZ) and scarps
seaward of the bank. .......................................................................................................................................................................................................................................................................................................................................................................................................................................................................
82
Figure 2. Isopach map of Late Triassic and younger sedimentary rocks in the Georges Bank basin plus the locations of the two COST wells. Thickness in kilometers dots show where thickness was measured along the profiles and indicates the control. ........................................................................................
82
Figure 1. A strike line from the Georges Bank region....................................................................................................................................................................................................................................................................................................................................................................................................................
85
Figure 2. Bear Seamount, one of the New England Seamounts, presumed to have formed above a thermal plume in the mantle...............................................................................................................................................................................................................................................................................................
86
Figure 3. Line taken southwest of the Bear Seamount. ..................................................................................................................................................................................................................................................................................................................................................................................................................
87
Figure 1. Bathymetric map of the Newfoundland basin showing location of the NB seismic grid. Portions of seismic profiles NB-2 (Figure 4) and NB-4 (Figures 6 and 7) are highlighted with a bold line. Note the location of the Skua E-41 industry well on NB-2. The migrated portion of NB-4
(Figure 8) is shown as a dashed line. Inset map shows the region of the southeastern Grand Banks shelf/eastern Canadian passive continental margin. ..........................................................................................................................................................................................................................................................
89
Figure 2. Correlation between Skua E-41 lithologies and computer-generated velocity and reflectivity logs. These logs were processed using the UTIG VAX 11/780 computer equipped with a DISCO Wavelet Processing package. ...........................................................................................................................
90
Figure 2a. Generalized lithologies derived from well-cuttings and sidewall core analyses. Ages are taken from paleontological studies conducted by AMOCO. (For the key to lithologic symbols, see Figure 4)..............................................................................................................................................................
90
Figure 2b. Velocity log converted from a hand-digitized sonic log. (This log was corrected to a sea surface datum by using velocities available from a check shot survey of the Skua E-41). ..............................................................................................................................................................................................
90
Figure 2c. Normal incidence reflectivity log generated from acoustic impedance values calculated from the velocity (Figure 2b) and density logs. ................................................................................................................................................................................................................................................................
90
Figure 3. Normal incidence travel-times and velocities for the Skua E-41 lithologic section shown on Figure 2 derived from a downhole check shot (normal incidence velocity/depth) survey. These were used both to guide the picking of stacking velocities for NB-2 and to determine equivalent two-way
travel times for major reflecting horizons after corrections were made for Conrads acquisition geometry (see also Figure 4). Velocities from the check shot survey are slightly higher than the NB-2 stacking velocities, but show a general correlation of major velocity breaks. .....................................................
90
Figure 4. Time section of the part of NB-2 that straddles the Skua E-41 well. (Actual closest-point-of-approach of the seismic line to the well location is 1.1 n. mi/6056 ft). Lithologies/ages are tied to the seismic section using the relationships illustrated on Figures 2 and 3. A high-amplitude, continuous
reflector at 1.15 sec correlates with the top of an Upper Cretaceous limestone section, and also ties to a large increase in velocity (Figure 2b) and a pronounced reflectivity spike (Figure 2c). However, this lithologic/acoustic boundary is too young to be the late Early Cretaceous U unconformity on the Grand
Banks. Another, less prominent reflector occurs at 1.29 sec and marks both a sandstone/limestone contact and a presumed Late Early Cretaceous unconformity. We correlate this geologic boundary with the regionally important U unconformity, which we can then successfully trace into the deep
northern Newfoundland basin. .......................................................................................................................................................................................................................................................................................................................................................................................................................................................
91
Figure 5. Magnetic anomalies plotted along dip lines of the NB survey. Shaded area is positive. Approximate trend of the J magnetic anomaly in the northern Newfoundland basin is highlighted. The J-anomaly is also plotted on Figures 6-8 above seismic line NB-4. ..............................................................
92
Figure 6. Newfoundland basin (uninterpreted time sectionNB-4). ............................................................................................................................................................................................................................................................................................................................................................................................
93
Figure 7. Newfoundland basin (interpreted time sectionNB-4). ................................................................................................................................................................................................................................................................................................................................................................................................
94
Figure 8. Newfoundland basin (migrated time sectionESE part of NB-4).................................................................................................................................................................................................................................................................................................................................................................................
95
Figure 1. Location map of the seismic lines; insert = location of the study area. ..........................................................................................................................................................................................................................................................................................................................................................................
97
Figure 2. Paleogeographic map of northwestern Great Bahama Bank at mid-Tertiary (?) time, showing the two north-south-trending depressions of the Straits of Andros and Bimini Embayment. Soundings in msec; stipple = 250-300 msec. (From Eberli and Ginsburg, 1987). ...................................................
97

Figure 3. WESTERN line: profile from Andros Island to the Straits of Florida showing complex fill of the Straits of Andros separating Andros Bank from Bimini Bank and the westward-prograding margin of Bimini Bank. (See Figure 1 for location.) (From Eberli and Ginsburg, 1987.) .................................
99
Figure 4a. Straits of Andros, a seismic profile of a section of Western line in Figure 3. See Figure 1 for location. ..................................................................................................................................................................................................................................................................................................................... 100
Figure 4b. Straits of Andros, interpretation of a section of Western line. The infill consists of an aggrading system followed by a high-energy, prograding system. See Figure 1 for location. ............................................................................................................................................................................................. 101
Figure 5a. Bimini Embayment, seismic profile (line 1-N-C). See Figure 1 for location. .............................................................................................................................................................................................................................................................................................................................................................. 102
Figure 5b. Bimini Embayment, interpretation of line 1-N-C: a low-energy, prograding system. .................................................................................................................................................................................................................................................................................................................................................. 103
Figure 1. Index map showing bathymetry near the Great Bahama Canyon and shot point locations for the seismic lines discussed in this chapter. Bathymetry contour interval in meters. ................................................................................................................................................................................................... 105
Figure 2. Extract of simplified geological map of the western Alps (after Gidon and Richard).................................................................................................................................................................................................................................................................................................................................................... 105
Figure 3A. Geoseismic section of profile BAC E12 part (3).......................................................................................................................................................................................................................................................................................................................................................................................................... 106
Figure 3B. Geoseismic section of profile BAC E12 part (2).......................................................................................................................................................................................................................................................................................................................................................................................................... 107
Figure 3C. Uninterpreted seismic profile BAC E12 part (1). See Figure 3E. ................................................................................................................................................................................................................................................................................................................................................................................ 108
Figure 3D. Interpreted geoseismic section and seismic profile BAC E12 part (1)......................................................................................................................................................................................................................................................................................................................................................................... 109
Figure 3E. Uninterpreted seismic profile BAC E12 part (2). See Figures 3B, 3F.......................................................................................................................................................................................................................................................................................................................................................................... 110
Figure 3F. Interpreted seismic profile BAC E12 part (2). .............................................................................................................................................................................................................................................................................................................................................................................................................. 111
Figure 3G. Uninterpreted seismic profile BAC E12 part (3). See Figures 3A, 3H. ....................................................................................................................................................................................................................................................................................................................................................................... 112
Figure 3H. Interpreted seismic profile BAC E12 part (3). ............................................................................................................................................................................................................................................................................................................................................................................................................ 113
Figure 4A. Uninterpreted seismic profile BAC E14 part (1). See Figure 4B................................................................................................................................................................................................................................................................................................................................................................................. 114
Figure 4B. Interpreted seismic profile BAC E14 part (1)............................................................................................................................................................................................................................................................................................................................................................................................................... 115
Figure 4C. Uninterpreted seismic profile BAC E14 part (2). See Figure 4D................................................................................................................................................................................................................................................................................................................................................................................. 116
Figure 4D. Interpreted seismic profile BAC E14 part (2). ............................................................................................................................................................................................................................................................................................................................................................................................................. 117
Figure 5A. Uninterpreted seismic profile BAC E8. See Figure 5B................................................................................................................................................................................................................................................................................................................................................................................................ 118
Figure 5B. Interpreted seismic profile BAC E8. ............................................................................................................................................................................................................................................................................................................................................................................................................................ 119
Figure 6A. Uninterpreted seismic profile BAC E15 part (1). See Figure 6D................................................................................................................................................................................................................................................................................................................................................................................. 120
Figure 6B. Uninterpreted seismic profile BAC E15 part (2). See Figure 6E. ................................................................................................................................................................................................................................................................................................................................................................................ 121
Figure 6C. Uninterpreted seismic profile BAC E15 part (3). See Figure 6F.................................................................................................................................................................................................................................................................................................................................................................................. 121
Figure 6D. Interpreted seismic profile BAC E15 part (1). ............................................................................................................................................................................................................................................................................................................................................................................................................. 122
Figure 6E. Interpreted seismic profile BAC E15 part (2)............................................................................................................................................................................................................................................................................................................................................................................................................... 123
Figure 6F. Interpreted seismic profile BAC E15 part (3). .............................................................................................................................................................................................................................................................................................................................................................................................................. 123
Figure 7A. Uninterpreted seismic profile BAC E9 parts (1) and (2). See Figure 7B..................................................................................................................................................................................................................................................................................................................................................................... 124
Figure 7B. Interpreted seismic profile BAC E9 parts (1) and (2)................................................................................................................................................................................................................................................................................................................................................................................................... 125
Figure 8A. Uninterpreted seismic profile BAC E5 part (1). See Figure 8C................................................................................................................................................................................................................................................................................................................................................................................... 126
Figure 8B. Uninterpreted seismic profile BAC E5 part (2). See Figure 8D................................................................................................................................................................................................................................................................................................................................................................................... 127
Figure 8C. Interpreted seismic profile BAC E5 part (1)................................................................................................................................................................................................................................................................................................................................................................................................................. 128
Figure 8D. Interpreted seismic profile BAC E5 part (2). ............................................................................................................................................................................................................................................................................................................................................................................................................... 129
Figure 9A. Uninterpreted seismic profile BAC E7. See Figure 9B................................................................................................................................................................................................................................................................................................................................................................................................ 130
Figure 9B. Interpreted seismic profile BAC E7. ............................................................................................................................................................................................................................................................................................................................................................................................................................ 130
Figure 10. Simplified geologic map of western Alps, in southeastern France. .............................................................................................................................................................................................................................................................................................................................................................................. 131
Figure 1. Locations of profile TD4 section, submersible dive sites (sites A, B, and C) and bathymetry of continental margin of the southeastern United States.............................................................................................................................................................................................................................................. 140
Figure 2. Detailed bathymetry of Blake Escarpment and location of section of seismic profile TD4 discussed here (from Gilbert and Dillon, 1981). This bathymetry is somewhat different from other published information (and from Figure 1), and we consider it more accurate than other bathymetric maps in the
vicinity of profile TD4.................................................................................................................................................................................................................................................................................................................................................................................................................................................................... 141
Figure 3. Model study of affect of Blake Escarpment velocity structure on seismic data. ............................................................................................................................................................................................................................................................................................................................................................ 141

xi

Figure 4. Ages of rocks sampled on the Blake Escarpment from ALVIN (Sites A, B, and C) and stratigraphic information from drill sites on the Blake Spur (see Figure 1 for locations). ................................................................................................................................................................................................... 142
Figure 5. Sketch of a section of the Blake Escarpment at profile TD4, based on observations from submersible ALVIN. See Figure 8 for location of dive observations relative to seismic profile....................................................................................................................................................................................... 142
Figure 6. Photograph of escarpment at 3892 m at site C, showing jointed stepped nature of the cliff. ......................................................................................................................................................................................................................................................................................................................................... 142
Figure 7. USGS profile TD4 Blake Escarpment time section. ....................................................................................................................................................................................................................................................................................................................................................................................................... 144
Figure 8. USGS profile TD4 Blake Escarpment interpreted time section...................................................................................................................................................................................................................................................................................................................................................................................... 145
Figure 9. Section of profile TD4 with migration applied (time section). ....................................................................................................................................................................................................................................................................................................................................................................................... 146
Figure 10. Section of profile TD4 with migration and depth-conversion applied.......................................................................................................................................................................................................................................................................................................................................................................... 147
Figure 1. Map of the eastern Gulf of Mexico, showing locations of seismic lines, dredges (Freeman-Lynde, 1983), ALVIN dives (Paull et al., 1984), and DSDP sites 535 and 540 (Buffler et al., 1984). .......................................................................................................................................................................... 153
Figure 2. Seismic lines AG-4 and SF-9 across the Early Cretaceous platform margin in the eastern Gulf of Mexico. Interpreted time sections. ...................................................................................................................................................................................................................................................................... 153
Figure 3. Seismic line GT2-11, both uninterpreted and interpreted time sections. The overall depth intervals of nearby ALVIN dive samples (2442-3266 m, Paull et al., 1984) and dredge samples (2000-3000 m, Freeman-Lynde, 1983) are superimposed on the escarpment. See text for a description of erosion
measurement. ................................................................................................................................................................................................................................................................................................................................................................................................................................................................................. 154
Figure 4. Seismic line GT2-11, both uninterpreted and interpreted time migrated sections. The overall depth intervals of nearby ALVIN dive samples (2442-3266 m, Paull et al., 1984) and dredge samples (2000-3000 m, Freeman-Lynde, 1983) are superimposed on the escarpment. See text for a description of
erosion measurement. .................................................................................................................................................................................................................................................................................................................................................................................................................................................................... 155
Figure 5. Seismic line GT2-24, both uninterpreted and interpreted time sections. The overall depth intervals of nearby ALVIN dive samples (2211-2925 m, Paull et al., 1984) and dredge samples (2200-3200 m, Freeman-Lynde, 1983) are superimposed on the escarpment. See text for a description of erosion
measurement. ................................................................................................................................................................................................................................................................................................................................................................................................................................................................................. 156
Figure 6. Seismic line GT2-24, both uninterpreted and interpreted time migrated sections. The overall depth intervals of nearby ALVIN dives samples (2211-2925 m, Paull et al., 1984) and dredge samples (2200-3200 m, Freeman-Lynde, 1983) are superimposed on the escarpment. See text for a description of
erosion measurement. ..................................................................................................................................................................................................................................................................................................................................................................................................................................................................... 157
Figure 1. Seismic line shot in platform environment, offshore western Florida, using a conventional horizontal scale. .............................................................................................................................................................................................................................................................................................................. 160
Figure 2. A squeezed (1:6 compression) seismic line, shot offshore western Florida. .................................................................................................................................................................................................................................................................................................................................................................. 161
Figure 1. Base map showing location of study area and data base. ............................................................................................................................................................................................................................................................................................................................................................................................... 165
Figure 2. Global cycle chart for interval of interest. From Vail et al., (1977, 1984), Todd and Mitchum (1979), and Mitchum and Uliana (1985). ................................................................................................................................................................................................................................................................... 165
Figure 3a. Seismic line 71, uninterpreted....................................................................................................................................................................................................................................................................................................................................................................................................................................... 166
Figure 3b. Seismic line 71. Note mounted units interleaved with the toes of prograding wedges, and the sands associated with these units in the well............................................................................................................................................................................................................................................................ 167
Figure 4a. Seismic line 109, uninterpreted..................................................................................................................................................................................................................................................................................................................................................................................................................................... 168
Figure 4b. Seismic line 109. Note the absence of the mounded lowstand fans in the basin and the sands located on the shelf in the well.................................................................................................................................................................................................................................................................................. 169
Figure 5a. Seismic line 72, uninterpreted....................................................................................................................................................................................................................................................................................................................................................................................................................................... 170
Figure 5b. Seismic line 72, showing a canyon cut into the shelf by the 132 Ma. unconformity. ................................................................................................................................................................................................................................................................................................................................................... 171
Figure 6. Correlated horizons are tied to the seismic interpretation and schematically represent the geometries seen on that data set........................................................................................................................................................................................................................................................................................ 172
Figure 7. Depositional model showing the relationship of subsidence and sea level to stratal patterns, facies, and lithology (from Vail, 1987)......................................................................................................................................................................................................................................................................... 173
Figure 1. Structural framework of Foz do Amazonas area (after Carozzi, 1981). ......................................................................................................................................................................................................................................................................................................................................................................... 175
Figure 2. Schematic geological section across the continental margin of northeastern Brazil. ..................................................................................................................................................................................................................................................................................................................................................... 175
Figure 3. Generalized stratigraphic section chart of Foz do Amazonas basin................................................................................................................................................................................................................................................................................................................................................................................ 175
Figure 4. Seismic location map of lines depicted in Figures 5 and 6, and including well 1-APS-33E.......................................................................................................................................................................................................................................................................................................................................... 175
Figure 5A. Uninterpreted seismic profile, northwest to southeast across Foz do Amazonas area, offshore Brazil. Well log for well 1-APS-33E correlates center of line................................................................................................................................................................................................................................ 176
Figure 5B. Interpreted seismic line of Figure 5A. Note well 1-APS-33E in center of line and in the index inset. ....................................................................................................................................................................................................................................................................................................................... 177
Figure 6A. Uninterpreted seismic profile south-southwest to north-northeast across Foz Do Amazonas area. See index map (Figure 4) for location of line and proximity to well 1-APS-33E............................................................................................................................................................................................. 178
Figure 6B. Interpreted seismic line of same section in Figure 6A. ................................................................................................................................................................................................................................................................................................................................................................................................ 178
Figure 6C. Geological depth section across the Paleogene depositional carbonate margin. The Oligocene margin has migrated about 3 km oceanwards relative to the late Eocene platform edge....................................................................................................................................................................................... 178
Figure 1. Location map of Para-Maranhao basin, offshore Brazil, including an index showing profiles A (Figure 2) and B (Figure 3)............................................................................................................................................................................................................................................................................... 179
Figure 2A. Normal time section of profile A, from the Para-Maranhao basin, offshore Brazil. See Figure 1 for location. .......................................................................................................................................................................................................................................................................................................... 180

xii

Figure 2B. Interpreted migrated section from profile A (see Figure 2A), across the Para-Maranhao basin, offshore Brazil. See Figure 1 for location. .............................................................................................................................................................................................................................................................
Figure 3A. Normal time section of profile B across the Para-Maranhao basin, offshore Brazil. See Figure 1 for location..........................................................................................................................................................................................................................................................................................................
Figure 3B. Interpreted migrated section of profile B (see Figure 3A) across the Para-Maranhao basin, offshore Brazil. See Figure 1 for location....................................................................................................................................................................................................................................................................
Figure 1. Location map of Potiguar basin, offshore Brazil, showing locations of profiles A and B (Figures 2 and 3). ................................................................................................................................................................................................................................................................................................................
Figure 2A. Time section of profile A, Potiguar basin, offshore Brazil. See Figure 1 for location.................................................................................................................................................................................................................................................................................................................................................
Figure 2B. Interpreted time section of profile A (Figure 2A), Potiguar basin, offshore Brazil. See Figure 1 for location............................................................................................................................................................................................................................................................................................................
Figure 3A. Time section of Profile B, Potiguar basin, offshore Brazil. See Figure 1 for location.................................................................................................................................................................................................................................................................................................................................................
Figure 3B. Interpreted time section of profile B (Figure 3A), Potiguar basin, offshore Brazil. See Figure 1 for location............................................................................................................................................................................................................................................................................................................
Figure 1.Location map showing the area of the Jequitinhonha basin, offshore Brazil, as well as the orientation of the seismic profile in Figure 2. ..................................................................................................................................................................................................................................................................
Figure 2A. Migrated section of seismic profile across the area of the Jequitinhonha basin, offshore Brazil. See Figure 1 for location. .....................................................................................................................................................................................................................................................................................
Figure 2B. Interpreted migrated section of seismic profile (Figure 2A) across the area of the Jequitinhonha basin, offshore Brazil. See Figure 1 for location.................................................................................................................................................................................................................................................
Figure 1. Location map showing the names (Italian and German) and locations of the most important Triassic carbonate buildups and platforms of the Dolomites. ......................................................................................................................................................................................................................................
Figure 2. Simplified stratigraphy of the western Dolomites. .........................................................................................................................................................................................................................................................................................................................................................................................................
Figure 3. The Ladinian model of platform evolution. (A) The theoretical scheme of aggradation versus progradation. (B) Observed relationships at the platform base. The left diagram depicts the relationship at a larger scale than the right diagram. The right diagram indexes an inset depicted in
Figure 3-C. (C) Observed relationships at the platform top. The left diagram depicts the relationship at a larger scale than the right diagram. The right diagram indexes an inset depicted in Figures 3-D..........................................................................................................................................................................
Figure 3-A. The horizontal progradation of the Catinaccio platform (about 6 km). Photograph taken from the top of the Latemar. Observe that the basal condensed successionthe Livinallongo Formation (Li)is thickening basinward. ................................................................................................................
Figure 3-B. Line drawing of Figure 3-A highlighting clinoforms. Index inset highlights area of photo enlargement used in Figure 3-C, depicting relationships at the platform base, discussed in Figure 3. .......................................................................................................................................................................
Figure 3-C. Close-up photo enlargement and subsequent line drawing of area highlighted in Figure 3-B. ..................................................................................................................................................................................................................................................................................................................................
Figure 3-D. The offlap relationship at the northern edge of the Latemar buildup. The horizontally bedded platform limestone progrades over the grossly clinostratified flank deposits. The relationship is depicted in general terms in the model shown in Figure 3. ..........................................................................
Figure 4. The Carnian model of platform evolution. (A) Theoretical model of progradation, using a Ladinian-type platform as the nucleus. (B) Observed relationships at the base (inset refers to photo in Figure 4-A) and at the top (inset refers to photo in Figure 4-C). (C) Diagrams depicting detail
of grafting the Carnian-type platform onto a Ladinian-type platform. .......................................................................................................................................................................................................................................................................................................................................................................................
Figure 4-A. Photograph showing the westward climbing progradation of the Sella platform.......................................................................................................................................................................................................................................................................................................................................................
Figure 4-A (1). Close-up photograph of Figure 4-A showing clinoforms detailed in Figure 4. ....................................................................................................................................................................................................................................................................................................................................................
Figure 4-A (2). Line drawing of Figure 4-A, including unit identifiers: SC, San Cassian Formation; R, Raibl Formation; DP, Dolomia Principale..................................................................................................................................................................................................................................................................
Figure 4-B. Panoramic view of the northern margin of the Sella platform. ...................................................................................................................................................................................................................................................................................................................................................................................
Figure 4-B (1). Line drawing of Figure 4-B, highlighting clinoforms in the Carnian-type platform. Insets show index to close-up areas of Figures 4-C and 4-D............................................................................................................................................................................................................................................
Figure 4-B (2). Geological profile across the Gardena Pass. LTr, Lower Triassic; L-Cb, Ladino-Carnian basinal succession; Lc, Ladinian core; R, Raibl Formation; DP, Dolomia Principale. ............................................................................................................................................................................................
Figure 4-C. The eastern wall of the Mesdi Valley (note inset in Figure 4-B (1) for location), with the clear oblique-parallel progradation pattern over a very thin San Cassiano Formation (the wooded ledge). Arrow at top indicates the thin, horizontally bedded toplap unit. ..........................................................
Figure 4-D. The same oblique-parallel progradation with horizontal toplap, at Gardena Pass. See Figure 4-B (1) inset for location..........................................................................................................................................................................................................................................................................................
Figure 4-E. The spectacular oblique-tangential prograding pattern of the Lagazuoi platform (Cunturines-La Varella Group, east of the village of San Cassiano). The horizontally bedded San Cassiano Formation (SC) is largely represented by carbonate turbidites and mudstone. Photograph at right
shows a distant view of the same outcrop. .....................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 5. Two classic stratigraphic relationships documenting the cessation of platform progradation: (above) fossilization by volcanics, and (below) fossilization by carbonate deposition. ..............................................................................................................................................................................................
Figure 5-A. The flank of the San Lucano-Civetta platform. The original paleoslope as well as the adjacent basin bottom have been suddenly fossilized by an enormous quantity of volcanic rocks (pillow lava, turbiditic hyaloclastite, and volcanogenic conglomerate) highlighted in the photo (V).
Because the area is only slightly deformed, it is actually possible to see the original basin depth (about 800 m). ...................................................................................................................................................................................................................................................................................................................
Figure 5-B. Onlap relationship of the Durrenstein Formation against the massive flank of the Colli Alti platform (San Vigilio Valley, at Mount Sadalce) demonstrates fossilization by carbonate burial. There is a structural dip of about 20 ( to the right. ...........................................................................................
Figure 1. Neuquen basin location map. ..........................................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 2. Regional stratigraphy of Neuquen basin. ........................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 3. Tithonian-Valanginian lithostratigraphic units................................................................................................................................................................................................................................................................................................................................................................................................................
Figure 4. Regional seismic cross section of Tithonian, Berriasian, and Valanginian sequences plotted in depth from seismic section along line A-A (see Figure 1). Generalized lithologies from wells are tied to section. ...............................................................................................................................................
Figure 5. Regional geologic cross section of Tithonian, Berriasian, and Valanginian stages, using the top of Valanginian as datum. Seismic sequence boundaries are correlated in wells. Wells are located along B-B (see Figure 1)..............................................................................................................................
Figure 6. Western half of regional seismic section located along line A-A. See Figure 1 for location. Uninterpreted section above, interpreted below. ...........................................................................................................................................................................................................................................................

181
182
183
185
186
187
188
189
191
192
193
195
195
196
196
196
197
197
198
198
199
199
200
200
200
201
201
202
203
203
203
206
206
206
207
207
208

xiii

Figure 7. Eastern half of regional seismic section located along line A-A. See Figure 1 for location. Uninterpreted section above, interpreted below. ............................................................................................................................................................................................................................................................
Figure 8. Correlation of Neuquen basin sequences with coastal onlap and eustatic sea-level chart (Haq, Hardenbol, and Vail, 1987); and correspondence of geometry of sequences with trends of eustatic changes. ........................................................................................................................................................
Figure 9. Seismic section along line C-C. See Figure 1 for location. ...........................................................................................................................................................................................................................................................................................................................................................................................
Figure 10. Depositional model of Tithonian-Valanginian depositional sequence..........................................................................................................................................................................................................................................................................................................................................................................
Figure 11. Shelf margins of Tithonian-Valanginian depositional sequences..................................................................................................................................................................................................................................................................................................................................................................................
Figure 12. Thickness map (contoured in two-way seismic time) and depositional environments of sequence B. ........................................................................................................................................................................................................................................................................................................................
Figure 13. Thickness map (contoured in two-way seismic time) and depositional environments of sequence F. .........................................................................................................................................................................................................................................................................................................................
Figure 1. Index map showing locations of profiles 601 and 603, discussed in this chapter. .........................................................................................................................................................................................................................................................................................................................................................
Figure 2. Biostratigraphy of DSDP leg 80. Left half represents the early Tertiary (Paleogene) zone in columns 548-551. The right half represents late Tertiary and Quaternary (Neogene) in same drill sites (Jar = Jaramillo event; Old = Olduvai event). Figure after Snyder et al., (1985). .....................................
Figure 3. Summary of sediment accumulation rates and distribution of unconformities for sites 548 to 551. A condensed interval is defined as having a sediment accumulation rate of less than 5m/m.y. Note that the only unconformities that occur across the entire margin are those in the middle to
upper Miocene, middle Oligocene, and upper Paleocene, although the middle part of the Upper Cretaceous is absent or highly condensed at all four sites (after Masson et al., 1985). .......................................................................................................................................................................................................

209
210
210
210
211
211
211
213
214

Interpreted seismic profile OC 601 parts (1) and (2), with accompanying geoseismic section for profile 601.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 601 parts (1) and (2).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 parts (3) and (4), with accompanying geoseismic section for profile 601.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 601 parts (3) and (4).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 parts (5) and (6), with accompanying geoseismic section for profile 601.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 601 parts (5) and (6).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 parts (7) and (8), with accompanying geoseismic section for profile 601.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 601 parts (7) and (8).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 parts (9) and (10), with accompanying geoseismic section for profile 601...........................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 601 parts (9) and (10)...............................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 parts (11) and (12), with accompanying geoseismic section for profile 601.........................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 601 part (13), with accompanying geoseismic section for profile 601.........................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (1) and (2), with accompanying geoseismic section for profile 603.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (1) and (2).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (3) and (4), with accompanying geoseismic section for profile 603.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (3) and (4).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (5) and (6), with accompanying geoseismic section for profile 603.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (5) and (6).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (7) and (8), with accompanying geoseismic section for profile 603.............................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (7) and (8).................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (9) and (10), with accompanying geoseismic section for profile 603...........................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (9) and (10)...............................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile OC 603 parts (11) and (12), with accompanying geoseismic section for profile 603.........................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile OC 603 parts (11) and (12).............................................................................................................................................................................................................................................................................................................................................................................................................

216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239

Figure 1. North-south line near the eastern margin of the basin. ...................................................................................................................................................................................................................................................................................................................................................................................................
Figure 2. North-south line parallel to that in Figure 1, and farther out in the basin.......................................................................................................................................................................................................................................................................................................................................................................
Figure 3. Part of the northern continuation of the line illustrated in Figure 2................................................................................................................................................................................................................................................................................................................................................................................
Figure 4. East-west line in eastern part of the basin.......................................................................................................................................................................................................................................................................................................................................................................................................................

241
242
243
244

214

xiv

Figure 5. East-west line in the center of the basin..........................................................................................................................................................................................................................................................................................................................................................................................................................


Figure 6. North-south line in the western part of the basin. ...........................................................................................................................................................................................................................................................................................................................................................................................................
Figure 1. Location of the profiles recorded in the Cap Ferret canyon. .......................................................................................................................................................................................................................................................................................................................................................................................
Figure 2. Location of outcrop photos A1-A7, Gres dAnnot region. .............................................................................................................................................................................................................................................................................................................................................................................................
Geoseismic section of seismic profile CF 110 parts (1) and (2 on right)..................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile CF 110 parts (1) and (2)..................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile CF 110 parts (1) and (2). .....................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile CF 106 part (3). ...................................................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile CF 106 parts (1) and (2). .....................................................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile CF 106 part (3)................................................................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile CF 106 parts (1) and (2)..................................................................................................................................................................................................................................................................................................................................................................................................................
Geoseismic section of seismic profile CF 109 parts (2) and (3). ...................................................................................................................................................................................................................................................................................................................................................................................................
Interpreted (above) and uninterpreted seismic profile CF 109 parts (1) and (2). ...........................................................................................................................................................................................................................................................................................................................................................................
Interpreted seismic profile CF 109 parts (1) and (2). .....................................................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile CF 109 parts (1) and (2)..................................................................................................................................................................................................................................................................................................................................................................................................................

245
246
248
248
249
250
251
252
253
254
255
256
257
258
259

Photo A1, Chalufy. .........................................................................................................................................................................................................................................................................................................................................................................................................................................................................


Photo A2, Chalufy. ........................................................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A3, Montagne de lAvalanche. .............................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A4, Mont Bertrand. ..............................................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A5, Tete Noire. .....................................................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A6, north of Chalufy.............................................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A7, northwest of Tete Noire. ................................................................................................................................................................................................................................................................................................................................................................................................................................................
Photo A8, Goleon. ..........................................................................................................................................................................................................................................................................................................................................................................................................................................................................

260
260
261
261
262
262
263
263

Index map of seismic lines. ............................................................................................................................................................................................................................................................................................................................................................................................................................................................


Figure 1. Dip line, oriented northwest-southeast, in the present-day slope-rise environment of Mozambique.............................................................................................................................................................................................................................................................................................................................
Figure 2. The delta plain built by the Zambesi River during the Late Tertiary. .............................................................................................................................................................................................................................................................................................................................................................................
Figure 3. A strike line in the delta, farther north along the same trend shown in Figure 2. ...........................................................................................................................................................................................................................................................................................................................................................
Figure 4. A strike line farther north than the preceding lines, away from the influence of deltaic sedimentation and possible delta switching. .........................................................................................................................................................................................................................................................................

265
266
267
268
269

Table 1. Main characteristics of the platform seismic facies units. ................................................................................................................................................................................................................................................................................................................................................................................................

270

Figure 1. Index showing location of seismic profiles Indus 10 and Indus 12, across the depositional fan of the Indus River. .....................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile Indus 10 parts (1) and (2)................................................................................................................................................................................................................................................................................................................................................................................................................
Geoseismic section of seismic profile Indus 10 parts (1) and (2). .................................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile Indus 12 parts (1) and (2)................................................................................................................................................................................................................................................................................................................................................................................................................
Uninterpreted seismic profile Indus 12 parts (3) and (4)................................................................................................................................................................................................................................................................................................................................................................................................................
Geoseismic section of seismic profile Indus 12 parts (1), (2), and (3)...........................................................................................................................................................................................................................................................................................................................................................................................

270
272
273
274
275
276

RIFT SYSTEMS

SYNRIFT AND POSTRIFT SEQUENCES IN


THE NORTHERN NORTH SEA

TOM PETTERSEN
PER TERJE SVELA
and
HALFDAN CARSTENS
Nopec Geo Services, A.S.
Trondheim/Stavanger, Norway
INTRODUCTION
This contribution describes briefly some structural and stratigraphic
elements of the development of the northern North Sea (58 to 62N). To
illustrate the evolution, four seismic segments (all taken from different
Nopec surveys) located in slightly different geological provinces will be
shown (Figure 1).

NORTH SEA RIFTING


The evolution of the northern North Sea took place during two
pronounced rift phases followed by postrift subsidence. The early rift phase
is of probable late Paleozoic age and can be demonstrated on seismic lines
both on the Horda Platform and farther south in the Stord basin (see Nopec
regional lines). The later rift phase of predominantly Late Jurassic age
easily can be demonstrated on seismic lines all over the northern North Sea,
including the Viking graben and the Tampen spur. The associated postrift
unconformity, generally known as the Base Cretaceous Unconformity, is a
pronounced seismic marker.
The rift phase, according to one model and elaborated upon by several
others, is characterized by crustal thinning in response to extension, and
concomitant subsidence through large-scale, listric, normal faulting. The
postrift phase involves thermal subsidence decreasing with time and
thickening of the lithosphere due to cooling. The late rift phase from basal
Callovian to basal Volgian spans a time period of some 30 m.y., while the
postrift phase, which is still in effect, has lasted some 135 m.y. The
resulting basin structure is characterized by tilted Jurassic fault blocks
separated by listric faults and buried by a thick Cretaceous-Tertiary infill.
Within the graben the prerift unconformity is at the top of the Middle
Jurassic Brent Group, which constitutes the main reservoir unit.

TAMPEN SPUR: SEISMIC LINE NOD 4-84-30


Introduction
This seismic line (Figures 5 and 6), running east to west in the
Norwegian sector of the North Sea, is part of a regional survey designed to
evaluate the Jurassic fault blocks so typical of the northern North Sea. The
purpose of this contribution is to analyze the seismic signatures of the
synrift and postrift sediments as they appear on one of these fault blocks.
The Viking graben of the northern North Sea is the extreme northern
continuation of the larger North Sea Rift (Figure 1). The rift is mostly
narrow, generally less than 60 to 70 km across, but north of 60N it widens
and includes the East Shetland basin. The Viking graben proper, however,
continues with a north-northeast strike. The seismic line analyzed is located
on the Tampen spur in the northeast prolongation of the East Shetland basin.
Figure 1. Index map showing four study areas described in this chapter.

T. Pettersen, P.T. Svela and H. Carstens

The Tampen spur is bounded to the east by the West Viking graben
boundary fault (Figure 2; from Karlsson, 1985) and to the northwest by the
South More basin boundary fault. The fault block discussed is clearly downfaulted
from another elevated fault block across most of the Snorre Escarpment.
Seismic Stratigraphy
The early rift phase sequences cannot be seen on this line. However, the
postrift subsidence (pre-later rift phase) is represented by a thick Triassic to
Middle Jurassic infill. The largely arenaceous continental deposits of the
Triassic give low reflectivity and thus few marker horizons. (Well control
proves substantial Triassic thicknesses.)
The base of the synrift sequence (prerift unconformity) is characterized
by marine onlap and is clearly identifiable in the back-basin of the tilted
fault block as top of the Brent reservoir. Below the eroded fault scarp east of
the crest, normal faults are present with the prerift unconformity partly
preserved in downfaulted minor blocks. Minor synthetic and antithetic faults can be
mapped along the out-of-the-basin dipping segment of the fault block.
An intra-synrift unconformity is interpreted as representing the base of
the marine, highly radioactive, anoxic Kimmeridge Clay shales (Draupne
Formation) and top of the marine Heather Formation. The Draupne
Formation is by far the most important source rock of the northern North Sea.
The postrift unconformity can be identified by marine onlap both in the
Viking graben toward the east and onto the eroded fault scarp, and in the
back-basin toward the west. Thus, during the early postrift phase in Early
Cretaceous time, marine clays progressively onlapped the tilted fault block
as the whole basin subsided in response to thermal relaxation. Planar
normal faulting developed and the crest of the structure retreated from the
main graben axis.

SOGN GRABEN EAST: SEISMIC LINE


NOD 4-84-19
Introduction
This seismic line (Figures 7 and 8), part of a large regional survey, runs
east to west in the eastern part of the Norwegian North Sea (see Figure 3).

Figure 2. Index map showing location of line NOD 4-84-30, Tampen Spur, and
relation to bounding features, northern North Sea.

Synrift and postrift sequences, northern North Sea

The geological section displayed on this profile focuses on the Upper


Jurassic section, which is the reservoir unit in the giant Troll field,
approximately 50 km further south. On the same trend, 50 km to the north,
exists a small condensate field, the Agat field, with a reservoir section of
Lower Cretaceous age.
The fault block shows an atypical North Sea fault-to-graben relationship
with the major fault being downthrown to the east away from the graben
axis. There is clear evidence that most of the movement of this fault took
place during Early Cretaceous time.
The whole Tertiary section, except an eastward-expanding Paleocene
unit, is truncated by the mid-Pliocene unconformity.
Seismic Stratigraphy
The early rift phase sequences cannot be defined on this line. The first
reliable seismic marker is the Middle Jurassic (top of the Brent) sequence
boundary. The parallel unit below this event most likely represents a
reduced Middle and Lower Jurassic section, consisting of interbedded,
shallow-marine, sand-shales at the base, and deltaic deposits at the top. The
deepest picked reflector probably represents a near top of the Triassic marker.
The Brent Group clearly is cut by numerous faults that show
displacements at least through the early part of Late Jurassic time.
The time equivalent of the reservoir section in the Troll field consisting
of three sandstone formations of Late Jurassic age (155 m.y. to 141 m.y.), is
found as a 300-msec-thick unit on the upthrown side of the fault block and
thins rapidly to the west, most likely as the result of erosion.
At least three individual depositional units can be identified inside the
megasequence (H-I). Of these, the lower two sequences have a
characteristic moundy and discontinuous internal-reflection pattern
suggesting a fairly high-energy environment during deposition. Between
this area and the Troll field there is a major east to west fault, believed to
have been active during deposition of these units. Sharing a common
sediment source with the Troll area, a possible northward-dipping slope
during deposition might explain the high-energy-environment character of
these potential reservoir rocks.
The megasequence subsequently eroded, particularly in the western
part, and was onlapped by Late Jurassic sediments (G-H), which are time-equivalent
to the prolific Draupne source rock found nearly everywhere in the North Sea.
The oldest part of this onlapping section (which can be mapped as a separate
sequence) exhibits, in part, a marked moundy character, which suggests a more sandy
unit than is normally found in the area for the same time period.
The mid-Cretaceous (F-E) sequence is found in the A gat field to

T. Pettersen, P.T. Svela and H. Carstens

the north and is ri ch in turbidite sands. The seismic ch a racter of the


i n t e rnal re flections seen on this line (stro n g, d i s c o n t i nu o u s , s l i g h t ly
m o u n dy or dow n l apping) stro n g ly suggests a turbidite section in
this area with the potential for good re s e rvoir ro ck s .

HORDA PLATFORM: SEISMIC LINE


NOD 3-84-39
Introduction
This line (Figures 9 and 10) was shot as part of a large semi-regional
survey in the Northern Viking graben and Horda platform offshore Norway.
The line is shot in an east-to-west direction over the Horda platform
between the giant Troll gas field to the west and the Norwegian mainland to
the east (see Figure 3).
Proven reservoirs in the area are the Lower Cretaceous turbidite
sandstones in the Agat field (120 km to the north), the Upper Jurassic
shallow marine sandstones in the Troll field (20 km to the west), and the
Middle Jurassic Brent sandstones and Lower Jurassic Statfjord sandstones
in various fields further to the west.

STRUCTURAL HISTORY AND SEISMIC


STRATIGRAPHY
Pre-Jurassic
The dominant feature in this section is the large Oygarden fault. This is
a basement-involved fault, situated at the margin of the Horda platform,
having a present-day displacement at basement level (M) of approximately
4000 m. Sediment deposition in the pre-Jurassic sequence is mainly
controlled by the tectonic movements of this fault. The faulting and
subsequent eastward tilting created a half-graben into which Paleozoic
sediments (L-M) probably were deposited in a setting similar to the
Hornelen basin 50 km to the northeast on the Norwegian mainland, where
deposition of the sandstones and conglomerates in Devonian alluvial fans
was controlled by a major fault at the basin margin.
Continued movement of the fault during the Triassic shed great volumes
of clastic material from the basement areas into the basin. A possible
alluvial fan development is recognized within the sequence K-L, rapidly
thinning westward away from the fault. Other fan developments can be
identified in the Triassic J-K sequence, with discontinuous reflections close
to the fault becoming more continuous in the mid- and distal-fan areas.

Jurassic and Post-Jurassic


The first sequence boundary that can be identified with any reliability is
the Lower Jurassic sequence boundary (I), representing the top of the
Statfjord Formation. This formation represents a transition from a
continental to a shallow-marine environment. The next sequence boundary
defines the top of the Brent Group (H), which is characterized by strong,
continuous reflections, possibly caused by areally widespread coal beds.
The Upper Jurassic interval, G-H, is time-equivalent to the three sandstone
formations that are the reservoir units in the Troll field. The sequence is
seen on the seismic line as a progradational unit, indicating a deltaic
environment, equivalent to the Troll field. Sediment supply was from the
east or northeast.
By the end of the Upper Jurassic, rising sea level led to deposition of the
organic-rich Draupne Formation shales (F-G). This is the principal source
rock in the area. The Lower Cretaceous sequence, E-F, onlaps the top of the
Draupne Formation (F). However, the erosion at the basin margin during the
Kimmerian phase was less severe than in the central parts of the basin.
Another unconformity surface, D, is within the Cretaceous megasequence,
cutting progressively deeper Cretaceous strata in a westward direction. This
means that the western part of this section was uplifted during the Late
Cretaceous period.
Associated with the Oygarden fault there is a rollover of the Jurassic and
Cretaceous section toward the fault. Across the other major fault in the
western part of the section, the Upper Cretaceous sequence, B-C, is at a
structurally higher position on the western side of the fault than on the
eastern. At the same time, the Upper Cretaceous beds show a slight rollover
toward the fault, whereas deeper beds are more straight near the fault. This
might suggest that strike-slip movements occurred, generating those
rollover features during the Laramide orogenic phase. Major uplift of the
Fennoscandian shield is marked by a major mid-Pliocene unconformity,
(A), cutting progressively deeper into Tertiary and Upper Cretaceous strata
in the landward direction.

EGERSUND SUB-BASIN: SEISMIC LINE CNST-82-O6


Introduction
Line CNST-82-06 (Figures 11 and 12) crosses the northernmost
prolongation of the Egersund subbasin in a northeast to southwest direction
(see Figure 4).
The Egersund subbasin is the northern extension of the Norwegian
Danish basin. The basin is bounded to the east by the Fennoscandian Border

Figure 3. Index map showing location of lines NOD 4-84-19, Sogn graben east,
and NOD 3-84-39, Horda Platform, and their relation to bounding features,
northern North Sea.

Synrift and postrift sequences, northern North Sea

Zone and to the west by a major north-to-south-trending fault that separates


it from the Sele High.
The Egersund subbasin started to develop during the Permian, with
deposition of marginal Zechstein evaporitic sequences. The basin continued
to subside until the Early Cretaceous. During Cretaceous and early Tertiary,
parts of the Egersund subbasin were subjected to inversion tectonics.
Particularly along the Fennoscandian Border Zone, inversion of relief
clearly can be seen. A thinning of the Upper Cretaceous chalk and erosional
topography in the central parts of the basin are also apparent.

REFERENCES
Badley, M.A., T. Egeberg, and O. Nipen, 1984, Development of rift basins
illustrated by the structural evolution of the Oseberg feature, block 30/6,
offshore Norway: Journal of The Geological Society, London, v. 141, p.
639-649.
Goff, J.C. 1983, Hydrocarbon generation and migration in the East Shetland basin
and Viking graben of the northern North Sea: Journal of The Geological
Society, London, v. 140, p. 445-474.
Karlssen, W., 1985, The Snorre, Statfjord and Gullfaks oilfields and the habitat of
hydrocarbons on the Tampen Spur, offshore Norway, in A.M. Spencer et
al., Habitat of Hydrocarbons on the Norwegian Continental Shelf,
Norwegian Petroleum Society, Graham and Trotman, p. 181-197.
McKenzie, D., 1978, Some remarks on the development of sedimentary basins:
Earth and Planetary Science Letters, v. 40, p. 25-32.

Seismic Stratigraphy
The Permian sequences cannot be seen on this line. The margin of the
Zechstein basin is located somewhat to the south.
The Triassic and Lower-to-Middle Jurassic sequences are all thinning to
the west. The internal configuration generally displays a divergent pattern
suggesting a synsedimentary activity along the Fennoscandian Border Zone
with alluvial fans building out from the fault scarps. Continental deposition
dominates the Triassic and Lower-to-Middle Jurassic sequences.
A relative rise in sea level during Late Jurassic time caused the shelf to
start building out from the Norwegian mainland toward the Egersund
subbasin. The internal reflection configuration displays a sigmoidal pattern
on this line, indicating a shale-prone deposit. An oblique pattern is observed
on lines located in the central part of the lobe.
The orientation of clinoforms seen in three dimensions gives an easterly
source for the lobe. It is tempting to believe that one of the northeast-tosouthwest-trending fjords on the southwest coast of Norway acted as a
source drainage area for the Upper Jurassic shelf. The base of this sequence
is most likely correlative to the mid-Kimmeridgian sequence boundary (141
m.y.), whereas the top is related to the base of the Valanginian sequence
boundary (133 m.y.).
Marine shales were deposited in an epeirogenic subsiding basin during
the Early Cretaceous, whereas the Upper Cretaceous sequence is dominated
by chalk. During Late Cretaceous-early Tertiary, inversion of relief took
place along the major Mesozoic faults. These movements are related to
strike-slip faulting along the Fennoscandian Border Zone. A regional uplift
of the northern part of the Egersund subbasin during Paleocene resulted in
the erosional topography that is seen on the seismic line at top of the
Cretaceous level.

T. Pettersen, P.T. Svela and H. Carstens

Figure 4. Index map showing location of line CNST 82-06, Egersund subbasin, and
its relation to bounding features, northern North Sea.

Synrift and postrift sequences, northern North Sea

STRUCTURAL AND STRATIGRAPHIC FRAMEWORK


OF THE JEANNE DARC BASIN, GRAND BANKS

H.J. WELSINK and A.J. TANKARD


Petro-Canada Resources
Calgary, Alberta
INTRODUCTION
The Grand Banks is the broad continental shelf that extends 450 km
offshore from Newfoundland. To the south and north it is separated from the
narrower shelves of Nova Scotia and Labrador by prominent transform
faults (Figure 1). Mesozoic extension resulted in a series of southwest-tonortheast-oriented basins. One of these, the Jeanne dArc basin, preserves at
least 14 km of Triassic through Lower Cretaceous sediments below the
Aptian break up unconformity. Since the discovery of the giant Hibernia
oil field, the Jeanne dArc basin has become the focus of Grand Banks
exploration, resulting in several other significant oil discoveries. The present
data base includes over 430,000 km of reflection seismic profiles and
observations from more than 80 deep exploration and delineation wells,
including 45 wells in the Jeanne dArc basin.
Two seismic profiles are used to illustrate here the structural and
stratigraphic evolution of the Jeanne dArc basin. Unconformity-bounded
sequences, which are the stratigraphic response to large-scale tectonic
processes, record major episodes in basin evolution. In the Jeanne dArc
basin they match the megacycles of rift and postrift subsidence as well as
various phases of ocean opening. Much of this paper has been extracted
from Tankard and Welsink (in press). However, it is our intention in this
study to highlight the nature of the unconformity-bounded sequences and
their relation to the structural evolution of the Grand Banks.

TECTONIC HISTORY AND GEOLOGICAL


SETTING OF THE GRAND BANKS
The continental margin of Atlantic Canada has a long history of
construction. The Paleozoic structural framework was built by the accretion
of numerous exotic terranes onto a late Precambrian-early Paleozoic passive
margin (Price and Hatcher, 1983; Williams, 1984). Late Paleozoic
reconstructions show the Grand Banks surrounded by the Canadian Shield,

H.J. Welsink and A. Tankard

and the African, Iberian, and European continental plates. Subsequent


extension in the Mesozoic was a complex process influenced by inherited
Paleozoic structures, sequential opening of the Atlantic Ocean, and major
reorganizations of oceanic plates (Tankard and Welsink, in press). On the
Grand Banks this resulted in a unique history of basin formation.
Mesozoic Extension
The Grand Banks records about 225 m.y. of basin evolution, including
two major episodes of rifting and postrift thermal subsidence (Figure 2).
The principal rifting event lasted about 50 m.y. from late Callovian to
Aptian, but appears to have inherited older structural trends of an aborted
Triassic rift system. Complicating this structural picture were several phases
of ocean opening in the mid-Jurassic, mid-Cretaceous, and end-Cretaceous
(Tankard and Welsink, in press). Figure 1 shows the principal tectonic
elements. Two major transfer faults (cf. Gibbs, 1984) divide the Grand
Banks into southern, central, and northern extensional terranes, each with
distinctive basin styles. Within the central Grand Banks we recognize three
major structural elements. Adjacent to Newfoundland, the Bonavista
Platform overlies thick unextended Avalon basement. To the east, late
Callovian-Aptian extensional tectonics produced a series of half-grabens,
including the Jeanne dArc, Carson, and Flemish Pass basins. The third
element is Flemish Cap, a large translated horst-block.

JEANNE DARC STRATIGRAPHY


The stratigraphic succession of the Jeanne dArc basin records four
basin-forming stages, including two episodes of rifting and postrift thermal
subsidence (Figure 2). Significant discoveries of hydrocarbons are restricted
to the late Callovian-Aptian synrift sediments. The principal reservoir
intervals are the Jeanne dArc, Hibernia, and Avalon sandstones.
Figure 1. The Grand Banks are the continental shelf offshore from Newfoundland. It is
separated from the Scotian and Labrador shelves by the Newfoundland and Charlie
Gibbs fracture zones, respectively. Internally, major strike-slip faults divide the Grand
Banks into southern (S), central (C), and northern (N) provinces. On a smaller scale,
transfer faults accommodate different amounts and rates of extension. Basin geometries
are based on seismic and gravity data. The ocean-continental boundary (OCB) is after
Masson and Miles (1984), the J-anomaly is after Sullivan and Keen (1978).

An early phase of Late Triassic-Early Jurassic extension lasted about 25


m.y., but was abandoned without the creation of oceanic crust.
Accumulation of continental redbeds, evaporites, and shales reflect the
gradual decay of rift intensity and encroachment of an epicontinental sea.
Subsequent regional thermal subsidence in Early to Middle Jurassic time
created a large shale-dominated epeiric basin.
Late Callovian-Aptian extension and subsidence resulted in
accumulation of thick sequences of alluvial and shallow marine sediments.
Early extension from late Callovian to Kimmeridgian is reflected in a thick
sequence of interbedded shallow-water limestones and shales, including the
prolific oil-prone source rock. The climax of rifting, 10 m.y. later in the
Kimmeridgian, and growing relief from major fault displacements, flooded
the basin with river-borne conglomerates and sandstones of the Jeanne
dArc Formation. In the Berriasian, finer grained sandstones were deposited
in the Hibernia region by prograding shoalwater delta systems. The
Hibernia reservoirs consist of stacked channel sandstones. After 20 m.y. of
intense rifting, subsidence rates decreased progressively, and overprinting
by thermal subsidence became more evident. From the Valanginian onward,
shorelines were persistently outside the confines of the Jeanne dArc basin,
resulting in deposition of major limestone markers (e.g., A and B markers).
The Avalon sandstones were deposited during the dying phase of the late
Callovian-Aptian rifting episode. These sandstones are fine-grained. Slower
subsidence along the Murre fault resulted in basinward progradation of
younger Avalon intervals.
Transition to the postrift era was characterized by the onset of transbasin
normal faulting (Figure 3) that reactivated older transfer fault trends. The
crests of rotated fault blocks preserve condensed sequences of reworked
sandstones (e.g., Upper Avalon sandstone). The Late Cretaceous-Tertiary
shelf was constructed by prograding shelf wedges and blanketing
limestones. In the Paleocene, prograding deltaic and submarine fans
converged on the Jeanne dArc depocenter. The modern continental terrace
wedge was built in Tertiary time, and is dominated by mudstones.

UNCONFORMITY-BOUNDED SEQUENCES
Rifting and ocean opening about the Grand Banks took approximately
120 m.y. to complete. Several milestones in basin evolution are recorded in

Structural and stratigraphic framework, Jeanne DArc basin

14

salt. Continuous reflection events in the lower part of the sequence are
attributed to Murre dolomites and limestones. Higher in the sequence, the
Whale limestone reflector is correlated with the Aalenian transgression. The
upper boundary unconformity is prominent along the eastern and southern
margins of the basin where uplift was substantial. Elsewhere, this Callovian
unconformity can be recognized as the transition to a new set of reflections
that belong to the late Callovian-Aptian rifting episode.
Several authors have equated the Whale limestone with separation of
Africa from Nova Scotia (Jansa and Wade, 1975; Given, 1977; Wade, 1981).
The coincidence with the Callovian unconformity of the Blake Spur
Magnetic Anomaly and spreading center shift south of the Grand Banks
also suggests a relationship between plate reorganization and extension on
the Grand Banks.
The Late Callovian-Aptian Sequence

Figure 3. Different structural styles have combined to dissect the Jeanne dArc basin
into a mosaic of small fault-bounded blocks. Two major fault trends are dominant:
southwest-to-northeast basin bounding faults; southeast-to-northwest cross-basin
faults, the major ones (thick lines) being transfer faults. A third trend exists on a
smaller scale: conjugate Riedel shears obliquely connecting transfer faults; potential
tear faults trending parallel or slightly obliquely to basin-bounding faults.

the unconformity-bounded sequences. Two seismic lines (Figures 5 and 7,


Profiles A and B respectively) show the seismic expression of the structural
and stratigraphic styles that resulted from the major tectonic processes.
Figure 4 shows a geological depth section of unconformity-bounded
sequences seen in Profile A (Figures 5 and 6).
The Late Triassic-Late Callovian Sequence
Figure 2. Tectono-stratigraphic column for the Jeanne dArc basin, showing major
episodes in basin evolution separated by prominent unconformities. These
unconformity-bounded sequences match the megacycles of rift and postrift
subsidence as well as various phases of ocean opening about the Grand Banks:
separation of Africa and Aalemian transgression (Whale limestone); Blake Spur
Magnetic Anomaly and spreading center shift (late Callovian extension); separation
of Grand Banks from Iberia (Aptian unconformity); separation of Flemish Cap from
Galicia margin (pre-Cenomanian unconformity); separation of the EuropeanGreenland plate (basal-Tertiary unconformity).

H.J. Welsink and A. Tankard

This sequence, spanning the first 65 m.y. of basin development, was


deposited during initial rifting and a subsequent period of thermal
subsidence. It is present in its entirety on Profile A (Figure 5). The lower
sequence boundary is an unconformity between Late Triassic red beds and
underlying Paleozoic basement. Its reflection is tenuous in places because
of limited seismic resolution at that depth, as well as the masking effect of

The rift megasequence spans approximately 50 m.y. and is bounded by


major late Callovian and Aptian unconformities. The succession is divided
into three smaller sequences, which are bounded by unconformities or
limestone markers. These units represent early, middle, and late stages of
rift subsidence.
The lower unit is bounded by the Callovian and Kimmeridgian
unconformities (Profile A; Figure 6). Internally, reflections are generally
continuous and of moderate amplitude, and result from limestone and shale
lithologies. The unit thins dramatically east (Figure 6, Profile A, between 6
and 8). The Kimmeridgian unconformity is also irregular and deeply
channeled along the eastern margin of the basin.
The middle unit is bracketed by the Kimmeridgian unconformity and
the B-marker, and shows a gradual change in reflection pattern upward
through the section. Variable amplitude, low-continuity reflection patterns
are associated with the Kimmeridgian unconformity; the pattern becomes
gradually more continuous upward (Figure 6; Profile A, between 3 and 6)
where coarse clastics grade into shales. The Hibernia sandstones in the
upper part of the sequence are difficult to resolve. Along the eastern ramp of
the basin, higher-frequency events (Figure 6, Profile A, between 6 and 8)
are caused by a condensed sequence of reworked sandstones and shales;
thinning of this interval resulted from continuous uplift of the margin of the
basin. The B-marker limestone, deposited during a regional transgression,
terminates the sequence.
The upper unit between the B-marker and the Aptian unconformity is a
relatively low-reflectivity unit that caps the late Callovian-Aptian
megasequence. Except in the Hibernia region, this unit is shaly with only
local occurrences of limestone stringers.
Erosion of salt-related structures (e.g., Hibernia) and elevated fault
blocks marked the onset of transbasin faulting (Figure 3). The Aptian

unconformity is the most prominent of a suite of unconformities that marks


the transition to a postrift era.
The Aptian-Albian Sequence
The pre-Cenomanian unconformity, forming the upper boundary of this
sequence, generally is masked by the strong reflection of the overlying
Petrel limestone or by its multiple. There also is interference with the
Aptian unconformity over the crests of some structural highs (e.g.,
Hibernia).
Several features are conspicuous. Synsedimentary faulting resulted in
variable stratigraphic thicknesses across the basin, with thickness variation
ranging from tens of meters to over 2 km (Figures 7 and 8; Profile B, 4
versus 10). Progressive rotation of fault blocks induced onlap onto structural
highs; these are expressed in condensed, reworked deposits (Figure 8;
Profile B, between 3 and 4). The mid-Cretaceous sequence was deposited in
predominantly marine environments in the transition to the postrift regime.
The pre-Cenomanian unconformity is correlated with separation of Flemish
Cap from the Galicia margin at about 100 m.y. (Masson and Miles, 1984).
The Late Cretaceous-Tertiary Sequence
The Petrel limestone is a continuous reflector, and introduces a new
style of basin development and evolution of the modern shelf. Prograding
shelf wedges (Figure 6; Profile A, between 2 and 6) show chaotic internal
reflections and deformation structures. Latest Cretaceous erosion formed a
prominent unconformity, on which the modern continental terrace wedge
was built. This basal-Tertiary unconformity coincides with the separation of
the European-Greenland plate, and northward tilt of the Jeanne dArc basin.

CONCLUSIONS
Several unconformities identified in the Jeanne dArc basin may
coincide with distinct phases of ocean opening about the Grand Banks.
Opening south of the Grand Banks, and the associated Blake Spur Magnetic
Anomaly (165 m.y.), resulted in the Callovian unconformity and the onset
of mid-Mesozoic extension on the Grand Banks. Separation from Iberia
(115 m.y.) is correlated with the major Aptian unconformity. Finally
separation from the Galicia margin (100 m.y.) is reflected in the preCenomanian unconformity. Opening of the Labrador Sea (80 m.y.) had no
equivalent unconformity in the Jeanne dArc basin. However, separation of
the European-Greenland plate (65 m.y.) might have been synonymous with
the unconformity at the base of the Tertiary.

Structural and stratigraphic framework, Jeanne DArc basin

15

ACKNOWLEDGMENTS
We thank Reed Johnson and G.S.I. for redisplaying the two seismic
lines. Drafting and typing were done by Terri Haber and Dawn Holmes.
Petro-Canadas support and permission to publish is much appreciated.

REFERENCES CITED
Gibbs, A.D., 1984, Structural evolution of extensional basin margins. Journal
Geological Society of London, v. 141, p. 609-620.
Given, M.M., 1977, Mesozoic and early Cenozoic geology of offshore Nova
Scotia. Bulletin Canadian Petroleum Geology, v. 25, p. 63-91.
Jansa, L.F., and J.A. Wade, 1975, Geology of the continental margin off Nova
Scotia and Newfoundland. Geological Survey Canada Paper 74-30, p. 51105.
Masson, D.G., and P.R. Miles, 1984, Mesozoic seafloor spreading between Iberia,
Europe, and North America. Marine Geology, v. 56, p. 279-287.
Price, R.A., and R.D. Hatcher, 1983, Tectonic significance of similarities in the
evolution of the Alabama-Pennsylvania Appalachians and the AlbertaBritish Colombia Canadian Cordillera. Geological Society America
Memoir 158, p. 149-160.
Sullivan, K.D., and C.E. Keen, 1978, On the nature of the crust in the vicinity of
the Newfoundland Ridge. Canadian Journal Earth Sciences, v. 15, p. 14621471.
Tankard, A.J., and H.J. Welsink, in press, Extensional tectonics and stratigraphy of
the Mesozoic Grand Banks of Newfoundland, in W. Manspeizer, ed.,
Triassic-Jurassic rifting and the opening of the Atlantic Ocean. Elsevier,
Amsterdam.
Wade, J.A., 1981, Geology of the Canadian Atlantic margin from Georges Bank to
the Grand Banks. Canadian Society Petroleum Geologists Memoir 7, p.
447-460.
Williams, H., 1984, Miogeoclines and suspect terranes of the CaledonianAppalachian orogentectonic patterns in the North Atlantic region.
Canadian Journal Earth Sciences, v.21, p. 887-901.

Figure 4. Geological depth section of Profile A (Figures 5 and 6) showing major unconformity-bounded sequences
and lithological units. Lithologies come from various wells projected on Profile A.

H.J. Welsink and A. Tankard

Structural and stratigraphic framework, Jeanne DArc basin

16

SEISMIC STRATIGRAPHIC ANALYSIS AND PETROLEUM EXPLORATION OF


PALEOGENE LACUSTRINE SANDSTONE BODIES, OFFSHORE BOHAI BASIN, CHINA

JIANG XI-JIANG
Verified by Liu Xingli
The Research Institute of Bohai Oil Company
Tang Gu, Peoples Republic of China
INTRODUCTION
During the early Tertiary, subduction of the Pacific plate in a westnorthwest direction and movement of the Indian plate in a northerly
direction caused both tension and right lateral shear in parts of the
continental crust of eastern China. A series of northeast-trending fault
blocks and grabens were formed in the Bohai basin area (Figure 1; see also
Li Desheng, 1981). Some of the grabens or sags contain rich Paleogene
lacustrine petroleum source rocks as well as sandstone, and locally
limestone, oil and gas reservoirs (Ma Li et al., 1982). In addition, oil
generated in these lower Tertiary source shales locally migrated into preTertiary reservoirs in buried hill traps. The onshore portion of the Bohai
basin has been explored and the major oil field complexes include Dagang,
Shengli, and Liaohoe. Although the offshore Bohai Bay part of the basin
has not been as fully explored, several commercial fields have been found.
Seismic data and drilling indicate that the center (or most subsident
parts) of the Bohai basin lie offshore under Bohai Bay where the Tertiary
strata are thickest (Figure 1) and where several types of Paleogene
sandstone facies are present. Some of these deposits have distinctive seismic
characteristics and may serve as models for exploration in other eastern China basins.
Unlike most marine environments, the Paleogene lakes of eastern China
were bounded by wrench-related, block-faulted, mountainous uplifts that
provided extensive and varied clastic source areas and drainage systems.
These sediments are arranged in predictable facies patterns controlled by
the stage of basin development and geologic process. During the early
block-faulting stage, more alluvial fans were developed because of the great
relief between grabens and horst blocks. During the middle stage, deep
lacustrine lakes filled the grabens and a number of short streams with steep
gradients provided sediments for subaqueous fans. During the late subsident
stage the fault depressions were characterized by fan deltas and deltaic

Jiang Xi-Jiang

deposits that filled in the lake depressions. In general, subaqueous fans and
fan deltas developed on the steeper, fault-bounded side of half grabens,
whereas smaller or thinner deltaic complexes developed on the more gentle
slope side. Locally, large-scale, high-relief, subaqueous prograding deltaic deposits
tended to fill the grabens along the long axis. Commonly, turbidite sands were
deposited along with prodelta mudstone in the central or deeper part of the lakes.
Seismic character of the different types of lacustrine and associated
deposits include external geometric form, internal reflection patterns,
reflection frequency, continuity and amplitude of the reflectors, and interval
velocity. Seismic stratigraphic analysis of three main types of sand bodies
associated with deep lacustrine environments in the Paleogene of the Bohai
basin are the subject of this chapter. Major structural features in the Bohai
Bay part of the Bohai basin, and the location of the seismic lines discussed,
are shown on Figure 1.

LACUSTRINE SUBAQUEOUS FAN DEPOSITS


Subaqueous fans, as used in this discussion, are sand and/or
conglomerate bodies formed by alluvial fans deposited partly within the
lake, and adjacent to a major uplift. Changes in sediment types being
supplied, topography, and the rise and fall of lake levels cause the
subaqueous fans to have different geometries and seismic reflection
character. Two examples are illustrated in this chapter.
Line1169.5 (Figure 2) shows the BN number 1 subaqueous fan located
in the HHK sag on the south side of the BN uplift. Figure 2A is an
uninterpreted migrated section with wavelet deconvolution processing. The
alluvial fan complex is interpreted and outlined on Figure 3A and overlies
basement. Its base is marked by a south-dipping, high-amplitude reflector
having good continuity and probably resulting from the smoothly eroded
basement surface. Its top is irregular and marked by a seismic facies change
from moderately continuous reflectors in the south, to short, discontinuous,
low-amplitude reflectors within the fan complex. Mapping the distribution
of this seismic facies indicates that the front and top of fan are characterized
by horizontal,more continuous, and moderate-amplitude reflectors which are, in turn,
onlapped and overlaid by parallel,continuous reflections within the lacustrine strata.
Based on depositional models of known subaqueous fans in the Bohai
basin and on velocity analysis, the lithologic characteristics and lithofacies

of the fan can be approximated. Figure 3B illustrates interval velocities and


the interpreted facies within the fan outlined on Figure 3A.
This subaqueous fan probably was deposited from a high-density
current adjacent to the mouth of a steep-gradient river and extended into a
lake characterized by deep-water lacustrine mudstone. Thus, potential
lacustrine oil source rocks interfinger with potential reservoir rocks. During
late stage regional subsidence, the fan was covered by dark mudstone of the
lower Dongying Formation (Ed), which forms the sealing cap rock. Well
No. 3 located on Figure 3D and the cross section of Figure 3E tested oil and
gas flows from the proximal part of this fan. The lithologic character of the
fan is shown on Figure 2C. It was deposited adjacent to an eroded uplifted
fault block (buried hill) having Cambro-Ordovician carbonate reservoirs
that tested commercial flows of oil and gas in Well 1 (BZ-28 field). Both the
fan and buried hill reservoirs may be connected to form a single field.
Figure 3E shows the relationship of wells 1 and 3 based on seismic line L 1,
which connects the two wells (Jiang and Zhao, 1984).
A second example of a subaqueous fan is shown on seismic line 1077,
located on Figure 1. The HZ Number 2 subaqueous fan is separated from
the basement by a fault plane forming the south flank of the HZ uplift. A
migrated dip section of this line is shown on Figure 4 and as shown on the
index map the wedge-shaped fan outlined on Figure 5 has higher amplitude
reflections at the top and base. The reflector dipping to the south is the fault
plane. Its poor continuity may result from erosional (?) topography along
the fault zone. Seismic facies within the fan include scattered low-amplitude
reflectors in the proximal part becoming more continuous in the middle
part, possibly because of increased shale beds. In the distal part of the fan
the reflection energy is locally weaker, possibly because of more frequent
mudstone beds. The top of the fan surface is distinct and is onlapped by
lacustrine mudstone. Compared to the BN number 1 fan (Figure 3), the
depositional velocity of the river sourcing the HZ number 2 fan probably
was faster, thus causing the front of the fan to spread farther into the lake.
Figure 5B shows the lithology of the proximal part of the HZ number 2
fan penetrated by the HZ number 4 well. The upper conglomeratic
sandstone is composed mostly of quartz and feldspar grains and the middle
part is mostly a granite pebble conglomerate. A schematic structural map at
the top of the fan surface is shown on Figure 5C. Figure 5D shows the

interpreted facies of the fan based on stratigraphic analysis of seismic line


1077. The HZ number 4 well was located on a rollover trap at the top fan
surface caused by continued movement of the fault after fan deposition. An
oil and gas flow was noted at the top of the fan in this well.
The subaqueous fans illustrated in Figures 3 and 5 appear wedge shaped
on seismic lines along depositional dip but are mounded or lenticular on
strike sections. In the Bohai basin, fans generally can be distinguished from
adjacent lacustrine strata, which are characterized by moderate to
continuous reflectors. The fans are believed to have formed by continuous
deposition from high-energy rivers having fixed river mouths. Apparently
the source rivers carried a heavy sediment load, possessed great density, and
the more distal deposits flowed along the lake bottom like a debris flow.
Where there was an abundant sediment supply and rapid sedimentation rate,
the proximal portion of the fan may have been partly above lake level.

FAN-DELTAS
Fan-delta, as used in this chapter, is an alluvial fan that progrades into
an ocean or lake environment. It is transitional between an alluvial fan and a
delta, and may be considered a laterally compressed deltaic complex. Some
workers believe most fan-deltas to be located at a continental margin and
formed by a high-gradient braided stream flowing into a depositional basin.
High relief topography produced by rifting also is a likely environment for
fan deltas if lakes are present in the grabens. Such was the case in the Bohai
basin depressions during the Paleogene.
During the Oligocene the paleo-Luanhe (river) fan delta prograded into
a deep-water lake on the south side of the SJT paleo-island (see Figure 1).
The drainage system originated in the Yanshan mountains north of the
Bohai Bay area. Migrated seismic line 7076 (Figure 6A) crosses the paleoLuanhe river valley cut into the basement north of the fan-delta. A highamplitude reflector marks the base of the river valley (Figure 7A). Possibly
the mounding within the valley-fill sequence is related to meandering of the river.
Migrated seismic section 67 (Figure 6B) is a northwest-to-southeast dip
section across the fan-delta outlined on Figure 7B. The border fault along
the south flank of the SJT uplift intersects the left-hand margin of the
section at 2.6 sec and forms a high-amplitude reflector dipping steeply to
the southeast. The wedge-shaped reflection package adjacent to the fault is

Main sand sedimentary bodies, Bohai Bay

22

interpreted to be a coarse clastic alluvial fan probably deposited near the


mouth of a canyon cut into the fault scarp. In front of this facies are at least
two sets of dipping, high-amplitude reflectors, which can be traced for a
long distance. They prograde southeastward, each eventually becoming
bottomsets and each having truncated tops. These dipping reflectors record
a moderately steep gradient into a deep lake. Possibly the strong reflectors
are due to acoustic impedance between sandstone and mudstone, and
suggest that large amounts of sand may have been transported down slope
into the deep lacustrine basin and deposited as turbidites. Velocity analysis
supports the interpretation that this prograding sequence is interbedded
sandstone and mudstone.
Figure 7D is our interpretation of the facies of this fan-delta. Fan-delta
sandstones locally are petroleum reservoirs in the Bohai basin because of
the concentration of oil-prone, mature organic matter in the enclosing
lacustrine mudstones. Near shotpoint 1600 on line 67 (Figure 7B), a
rollover anticline is present on the southeast side of a growth fault. Special
processing of the seismic data in this area indicates dimming of the top sandstone
reflector over the anticline, suggesting the possibility of hydrocarbons.
A fan-delta complex extending eastward into the LZW deep depression
(sag) is illustrated on west-east Line 134 of Figure B and is outlined on
Figure 9A between 1.8 and 3 sec (two-way time). Note the continuity of the
overlying and underlying lacustrine mudstone interpreted to have been
deposited in deep water. The source of sediment is the KD uplift bordering
the LZW sag on the west (see Figure 1). Downthrown to the western border
fault, the interpreted proximal fan facies is characterized by scattered,
chaotic, short, discontinuous reflections (stippled pattern). This seismic
facies grades eastward into a sequence of progradational reflectors that are
more steeply dipping in the upper part of the foresets and became gentle as
they grade into prodelta bottomsets. The foresets are younger and less
steeply dipping on the east part of the line, probably because of decreasing
current energy, shallowing of the lake because of rapid deposition, and/or
diapiric uplift of the eastern margin of the lake basin during deposition. In
the distal or prodelta segment, the wavy hummocky reflections between
shotpoints 400 and 280 of Figure 9A are interpreted to be turbidite deposits
composed of sand and clay. Figure 9B show the interpreted lithology and
facies within the prograding fan sequence.

LARGE-SCALE DELTAIC PROGRADATION


The sedimentary environments of lakes are somewhat different from
those of oceans. Lacustrine delta geometry and deposition are controlled
mostly by the action of rivers. The shape of lacustrine deltas is greatly
influenced by the topography of the lake boundaries, being elongate along
the long axis of half-graben depressions and more lobate along the sides of
the graben.
The largest Paleogene graben depression in the Bohai basin is the

Bozhong Depression (BZ) offshore Bohai Bay (Figure 1). During the
Oligocene this depression was the site of a deep-water lake into which
several deltaic complexes prograded during lower Dongying (Ed)
deposition. A major, large-scale deltaic sequence prograded southwest more
than 150 km into the Bozhong Depression along its long axis. A part of this
sequence is illustrated on Line 138 of Figures 10 and 11. It is a spectacular
example of a lacustrine deltaic complex prograding into deep water.
As shown on Figure 11, the prograding deltaic sequence intersects the
northeast end of Line 138 between 2 and 3 sec and between 3 and 4 sec on
the southwest end of the line. Regional study suggests that deltas
prograding perpendicular to the narrow LZ depression converged to provide
clastic sediments for a major axial delta that prograded southeast into a deep
lake occupying the Bozhong depression during the Oligocene. Time-depth
plots indicate that the vertical distance between equivalent topset and
bottomset reflections at several locations on this line is approximately 1000
m, and demonstrate that this part of the Bozhong lake was very deep. The
irregular base of the sequence may be due to erosion and block faulting of
the underlying Shahejie (Es) and older deposits. The deltaic facies
illustrated on Figure 11A probably include topset deltaic plain deposits,
deep slope deposits, as well as deep lacustrine bottomset muds and turbidite
sands.

CONCLUSIONS
The three types of sand bodies illustrated in Figures 2 through 11 are
developed in the Paleogene lacustrine depressions of the Bohai basin. Oil
and gas reservoirs are found in each of these facies onshore, and potentially
commercial petroleum accumulations are found in subaqueous fans in the
offshore Bohai Bay portion. The Niuzhuang delta in the onshore DY
depression is potentially a commercial petroleum reservoir and has seismic
characteristics very similar to the deltaic deposits illustrated in this chapter.
Therefore, the sand bodies described here are not only of theoretical
significance, but also may be important models for future exploration in
basins similar to the Bohai.

REFERENCES CITED
Jiang Xi-jiang, and Zhao Yao,1984,Analysis of the BZ 28-1 buried subwater alluvial fan:Oil
Geophysical Prospecting, v. 1, p. 87-93 (in Chinese).
Li Desheng, 1981, Geological structure and hydrocarbon occurrence of the Bohai
oil and gas basin (China), in Petroleum geology in China: Tulsa,
Oklahoma, Penn Well Books, p. 180-192.
Ma Li, Ge Taisheng, Zhao Xueping, Zie Taijunn, Ge Rong, and Dang Zlienrong,
1982, Oil basins and subtle traps in the eastern part of China, in M.T.
Halbouty, ed., The deliberate search for the subtle trap: AAPG Memoir 32,
p. 287-315.

Figure 1. Index map of major structural features


offshore, Bohai Bay, PRC, and location of seismic
lines; contour in km on the base of the Tertiary.

23

PASSIVE MARGINS

INTERPRETATION OF WEST AFRICA


LINE C

R. E. SHERIFF
University of Houston,
Houston, Texas
INTRODUCTION
The West Africa lines that Esso Production Research contributed to the
AAPG as examples of seismic stratigraphic interpretation probably have
been interpreted by more people trying to learn seismic stratigraphic
methods than any other data set. Line C has been used in almost all of the
AAPG Seismic Stratigraphy Schools as well as by students in university
classes.
An interpretation of this line is shown on page 157 of AAPG Memoir 26
(Todd and Mitchum, 1977) (Figure 1). More and more interpretative detail
has been discovered in this line insofar as it has been discussed in AAPG
schools over the years. In 1985, Esso Production Research made the
magnetic tapes for this line available to the AAPG, and Fred Hilterman of
the Geophysical Development Corporation offered to reprocess these data
for the AAPG, free of charge. Edip Baysal supervised the processing, which
resulted in three new outputs: (1) a reprocessed section plotted to the same
scale as the original line (Figure 3); (2) a print of the stacking velocities
used in the reprocessing; and (3) a depth section (Figure 11).
These results have been used in recent AAPG schools. My objective
here is to show some of the changes in interpretation that have resulted from
the reprocessing.

COMPARISON OF REPROCESSED AND


ORIGINAL SECTIONS
There are three major differences between the original and the
reprocessed sections:
1. The embedded wavelet (effective wavetrain from the source, and
changes in wavelet shape in passing through the Earth and instruments) was
shortened and converted to zero-phase by wavelet processing.
2. Amplitudes are preserved (the original processing used an automatic
gain control).

R. Sheriff

Figure 1. Interpretation of West Africa Line C (shown in AAPG Memoir 26, p. 157).

3. The reprocessed data are migrated (the original line was not
migrated).
The original section is shown in Figure 2 and the reprocessed section in
Figure 3; an interpretation of Figure 3 is shown in Figure 4. The wavelet
processing removed the very ringy character of the original line, made
discrete reflections stand out more clearly, and gave more character to the
section. The amplitude preservation subdued very weak events, which the
automatic gain control (AGC) action strengthened. The migration collapsed
diffractions and sharpened evidence of faults and other features. Because all
events on the section were nearly horizontal, the migration has not done
much to move features to different locations, so its effects were not great.

Most changes made interpretation easier. However, amplitude preservation


made it more difficult to see weak, top-lapping reflections in at least a
couple of locations (Figure 5).

CONCLUSIONS FROM THE VELOCITY DATA


The locations of the velocity analyses are shown in Figure 6, along with
interpolated contours of the stacking velocity superimposed on a print of the
section. It is clear from the three velocity analyses over the shelf at the right
end of the line (shot points 1505, 1425, and 1279, Figure 7) that the Jurassic
shelf sediments are principally carbonates (as they were originally

interpreted) because the interval velocities are simply larger than are
expected for a sand-shale section. The consistency among these three
analyses is very high, so the conclusions drawn from them appear reliable. I
believe that more stratigraphic detail can be discerned if more analyses are
run and interpreted in greater detail, but this has not been done.
The three velocity surveys over the intermediate shelf (shotpoints 1131,
992, and 810) are shown in Figure 8. These analyses scatter considerably;
more analyses might have clarified the reasons for the scatter and might
have shown detail that would have helped interpretation. Nevertheless, the
analyses show a major change in the velocity at the Jurassic shelf edge and
this changes the original interpretation in an important regard. Figure 1
interprets the Jurassic section as being thin over the intermediate shelf by
projecting the Top of the Triassic sequence boundary through the poor data
area under the Jurassic reef and shelf edge. However, the depth of the Top of
the Triassic over the right Jurassic shelf (at 2.9 sec) is about 19,600 ft (6000
m), which corresponds to an arrival time of about 3.4 sec under the
intermediate shelf. The velocity suggested by the three analyses is in the
14,000 to 15,000 ft/sec range (4267 to 4572 m/sec), which gives a Jurassic
thickness of 4600 to 4900 ft (1400 to 1490 m), compared to about 1500 ft
(460 m) as would be inferred from Figure 1. This is about the same as the
4800 ft (1460 m) indicated over the right Jurassic shelf.
Although poorly defined, the velocities of the Cretaceous rocks over the
intermediate shelf are slightly lower than those over the right Jurassic shelf
and just slightly larger than would be expected for normally-pressured
Tertiary clastic sediments subjected only to compaction because of the
weight of the overburden (the dotted curves shown in Figures 7, 8 and 9;
from Sheriff and Geldart, 1983, p. 8). This suggests that the Cretaceous
sediments over the intermediate shelf are finer grained. It also rules out the
possibility that the layers contain large amounts of carbonates, which might
be expected if they were derived from erosion of the adjacent carbonate
shelf.
Interval velocities for the six analyses at the left end of the line
(shotpoints 715, 699, 530, 350, 249, and 199), where the ocean begins to
deepen more rapidly, are shown in Figure 9a. The tremendous scatter of
these Figure 2. Original section used in AAPG Seismic Stratigraphy
Schools.analyses is a consequence of plotting them with respect to a sea-

Interpretation of West Africa, line C

37

level datum. Velocity within rocks is, in part, a consequence of the


differential pressure to which they are subjected. The addition of an
overlying column of water adds the water weight equally to the overburden
and fluid pressures, and thus does not affect the differential pressure. Before
drawing conclusions from the velocity data, one should reference them to
the sea floor, as was done in Figure 9b. This considerably reduces the
scatter. Much of the data indicate lower velocities than those expected for
normally pressured sediments, suggesting under compaction. The data also
show abnormally high fluid pressures, which in turn suggest very little
permeability along with fine-grained sediments, because otherwise the

Figure 7. Interval velocities calculated from the velocity analyses using the Dix
equation for the three surveys over the Jurassic shelf. The solid curve shows the
expected velocities for a Tertiary clastic section subjected only to the pressure of the
overburden (from Sheriff and Geldart, 1983, p. 8). See Figure 6 for location.

R. Sheriff

abnormal pressures probably would have leaked off. This appears to apply
to almost all of the Tertiary section as well as to parts of the Cretaceous
section.

FAULTING ASSOCIATED WITH SHALE FLOW


The Cretaceous section shows growth faults just seaward of both the
right Jurassic shelf and the intermediate shelf. These are expected because
of the marked changes in the foundation provided by the underlying
Jurassic rocks. The growth faults are probably facilitated by abnormal
pressures associated with low permeability in underlying fine-grained

Figure 8. Interval velocities for the three surveys over the intermediate shelf. See Figure 6
for locations.

Cretaceous shales. The velocity data are not definitive enough to make a
convincing case for this, but the interval velocity values lying below the
normal compaction curve hint at it. If the section seaward of the
intermediate shelf edge (Figure 10) were abnormally pressured shale, it
should flow to the left and then upward, which would be down the pressure
gradient at the end of Cretaceous time. Anticline-shaped reflections under
shotpoint 550 are interpreted as shale diapirs. The other folded structures in
the Cretaceous at the left end of the section are likewise interpreted as shale
flow.
The pre-Cretaceous probably is faulted on the slope in front of the

intermediate shelf, but these faults likely are an independent system from
the Cretaceous growth faults. The latter probably get lost in the semi-fluid
abnormally pressured shale in the Lower Cretaceous, rather than continue
into the pre-Cretaceous.
The relief on the present sea floor toward the left end of the line is
attributed to mass movement downslope (also shown in Figure 11).
Fragmentary reflections that are nearly horizontal line up with the sea-floor
relief features; these are interpreted as reflections marking the bases of
slumps. Reflections generally are parallel to the sea floor, but there are no
reflections above these flat reflections over a zone about 40 km wide. The

Figure 9. Interval velocities for six of the velocity surveys over the left end of the line
where the water is deepening. A. Plotted with respect to a sea-level datum; B. plotted
with respect to a sea-floor datum.

Interpretation of West Africa, line C

42

present sea-floor dip is about 0.9 degree and probably was never greater
than this, but massive downslope movement surely occurred. Velocity
analyses in this area indicate values only slightly greater than water
velocity. This is interpreted to indicate undercompacted fine-grained muds
containing excess interstitial water that cannot escape because of very low
permeability; the consequent lack of shear strength makes the sediments
behave as a fluid. Any bedding originally present was probably destroyed in
the downslope movement, thereby producing the nonreflection character.
The fact that the sediment velocities are nearly the same as those of
water also explains why the water-bottom multiple(s) are so prominent in
this area. Common-midpoint stacking was the only process applied to affect
multiple attenuation, and it is not effective unless multiples and primaries
have different stacking velocities.

ACKNOWLEDGMENTS
I thank Fred Hilterman and Edip Baysal of the Geophysical
Development Corporation for their reprocessing contributions, and Esso
Production Research and Peter Vail for providing the data.

REFERENCES CITED
Payton, C.E., ed., 1977, Seismic stratigraphyapplications to hydrocarbon
exploration: AAPG Memoir 26, 516 p.
Sheriff, R.E., and L.P. Geldart, 1983, Exploration seismology, volume 2:
Cambridge England, Cambridge University Press.

CONCLUSIONS FROM THE DEPTH SECTION

Figure 10. Interpretation of flow features at the left end of the line.

R. Sheriff

The same data are shown in Figure 11 plotted with a scale linear in
depth rather than linear with time. A number of features appear different as
consequences of this change. The relative thicknesses of various portions is
especially altered. The Jurassic shelf at the left end of the section is seen to
be much thicker and the Tertiary section at the right end much thinner. The
erroneous inferences about the thickness of the Jurassic section above the
intermediate shelf referred to earlier would not be as likely with the depth
section. The depth section is not as reliable as the time section because it
incorporates the additional uncertainties about the velocity. Hence, one
would not want to interpret depth sections without careful review against
the time sections, just as one would not want to interpret migrated sections
entirely without checking against the unmigrated sections. However, the
changed viewpoint that the depth section affords seems sufficient
justification for wanting to have it available in an interpretation.
Figure 11 also has only about a 2:1 vertical exaggeration, compared to
6:1 to 8:1 for Figure 3. The changed ratio also makes many features appear
different; the interpreter has a better perspective for the reef at the Jurassic
shelf edge and for the dip of faults. It is common to plot seismic sections for
stratigraphic interpretation with very large vertical exaggeration, which
greatly distorts structural relationships. The common view is that large
vertical exaggeration is necessary to reveal angularities in the data (which
are so important in stratigraphic interpretation). However, most angularities
can be seen with about equal ease on Figure 11 and on Figure 3, so perhaps
the common wisdom is challenged.

Interpretation of West Africa, line C

43

Figure 11. Display of the reprocessed section linear in depth rather than linear in time, and
with an aspect ratio (vertical to horizontal scale) of 2:1.

LISTRIC FAULTS, OFFSHORE MOROCCO

D. BRADFORD MACURDA, JR.


The Energists
Houston, Texas
The stratigraphic evolution of passive margins is strongly affected by
the nature of the early sediments formed on these margins. An unstable
substrate, such as shales or salt, which can undergo plastic deformation due
to loading, sets up many new exploration plays on siliciclastic margins
(Winker and Edwards, 1983). This is due to both syndepositional and
postdepositional structuring. If a margin has an early history of carbonate
deposition, this also can strongly influence the subsequent evolution of the
depositional systems. Carbonate margins often have an abrupt shelf-slopebasin transition because of the penecontemporaneous cementation of the
carbonates of the shelf margin. This produces a steep gradient with a greatly
expanded section downdip. If siliciclastic sediments are part of the
subsequence depositional systems, the profile is steeper than the angle of
repose of the siliciclastics. Gravitational failure of the infilling sediments
occurs, forming listric faults. This is well illustrated by parts of the African
and South American margins.
The seismic sequences and geological development of the continental
margin of Morocco were discussed by Hinz, Dostmann, and Tritsch (1982).
There is a prominent series of Tertiary slope deposits, designated as (1) in
Figure 1, in the Tarfaya segment of the margin; these deposits range in age
from Paleogene through Pleistocene, with many unconformities within
them. The water-bottom multiple (2) cuts across these slope deposits and
should not be confused with them. The upper part of the Cretaceous shelf
deposits (3) is erosionally truncated at the seaward edge and unconformably
overlain by the Tertiary slope deposits (1). The shelf deposits are
Hauterivian-Albian in age. The Late Jurassic carbonate shelf and margin (5)
and the post-Jurassic steep slope into the basin are very evident. The
expanded section (4), which infills the void, is the downdip equivalent of
the Cretaceous shelf facies. Because faulting was repetitive, including both
further rifting, which affected the Jurassic deposits, and gravitational
failure, the environments represented by the sediments in (4) probably
represent a complex interplay of shelf, slope, and basinal deposits. The

D.B. Macurda, Jr.

faulting of the Jurassic basinal deposits probably was induced by movement


of Jurassic salts, which lie 10 to 45 km seaward of the Jurassic shelf margin.
A second section (Figure 2), which lies a short distance to the southwest
of Figure 1, shows well the configuration of the listric faulting and the
offset of various units. The same numbers apply as in the discussion of
Figure 1. The seismic stratigraphic and facies analysis of such areas must be
carefully done because potential reservoirs tend to be rotated and offset by
the faulting. Winker and Edwards (1983, p. 141) give criteria for locating
shelf margins in unstable progradational systems. As they and many other
authors have noted, areas with listric faulting of the type illustrated herein
can have substantial petroleum reserves, as in the Gulf of Mexico.

REFERENCES CITED
Hinz, K., H. Dostmann, and J. Tritsch, 1982, The continental margin of Morocco:
seismic sequences, structural elements, and geological development, in U.
von Rad, K. Hinz, M. Sarthoin, and E. Siebold, Geology of the northwest
African margin: Berlin-Heidelberg, Springer-Verlag, p. 703.
Sheriff, R.E., 1988, Interpretation of West Africa Line C, this volume.
Winker, C.D., and M.B. Edwards, 1983, Unstable progradational clastic shelf
margins: SEPM Special Publication 33, p. 139-157.

Listric faults, offshore Morocco

45

AN UPDATE OF U.S. GEOLOGICAL SURVEY SEISMIC REFLECTION LINE 25


ACROSS THE NEW JERSEY SHELF, SLOPE, AND UPPER RISE

J.S. SCHLEE
C.W.POAG
U.S. Geological Survey
Woods Hole, Massachusetts
and
J.A. GROW U.S.
Geological Survey
Denver, Colorado
INTRODUCTION
In the three years since Grow et al. (1983) published an interpretation of
United States Geological Survey (U.S.G.S.) line 25 over the New Jersey
margin, at least two studies have been made in the vicinity of this line. In
1983, Deep Sea Drilling Project (DSDP) legs 93 and 95 drilled holes along
this line to date the post-Cretaceous continental slope and rise section
(Poag, in press). Second, in 1986, Schlee and Hinz (1987) completed a
study of the seismic stratigraphy of the slope and upper rise based on
U.S.G.S. profiles and a 1979 survey of this area by Bundesanstalt fur
Geowissenschaften und Rohstoffe (BGR). Later exploratory drilling that
continued into 1985 mainly covered the upper 15,000 ft (4.6 km) of
sedimentary section. Hence, data on reflectors 4-7 (uppermost Jurassic to
pre-Lower Middle Jurassic; Table 1) under the continental shelf and in the
deep sea are unchanged from what Grow et al. (1983) published, and their
analyses are used in this interpretation (Table 1). Though the older reflectors
are not reinterpreted, the profile itself has been reprocessed to evaluate the
deep structure beneath Baltimore Canyon Trough.

PRESENT PROFILE
Reprocessing involved changing the vertical and horizontal scales so
that line 25 could be shown on one profile instead of three (as was done by
Grow et al., 1983). The profile is displayed and interpreted in a depth
profile (Figure 1); vertical exaggeration is 5:1. The Grow et al. (1983)
article offers a complete discussion of the tectonism and basin development

J.S. Schlee, C.W. Poag and J.A. Grow

of the mid-Atlantic margin of the United States. It is not our purpose to


repeat what was well described in the original atlas, including the true
amplitude section, and gravity and magnetic models that were given along
the same line.
We focus on minor modifications to the Cenozoic part of the section on
the basis of data from the two later studies, but we strongly urge study of
the earlier Grow et al. (1983) paper and the Poag (1987) summary of drill
hole results; some of his results are used on Figure 1.
On the basis of a 2350-km multichannel, seismic-reflection survey of
the outer shelf, slope, and upper rise off northern New Jersey, Schlee and
Hinz (in press) divide the Cenozoic section into two main rise sequences
(and several sub-sequences). The sequences are separated by a conspicuous
unconformity inferred to be A[u], by analogy to other studies (Schlee et al,
1985). During the Cretaceous the slope-rise transition was much gentler and
the sedimentary sequences were more blanket-like and continuous from rise
to shelf.
On the basis of an examination of cores from DSDP holes on the slope
and upper rise area and 2350 km of the same multichannel profiles used by
Schlee and Hinz (in press), Poag (in press) mapped 12 major Upper
Cretaceous and Cenozoic sequences. Most sequences are documented in
five DSDP drill holes, where they are separated by unconformities that are
marked by indications of downslope mass-sediment displacement (turbidity
currents, debris flows, and slumps). The unconformities are widespread and
can be traced over most of the slope and rise area, where they correlate with
similar unconformities under the shelf and on the adjacent coastal plain
(Poag, 1987); in the erosionally complex area of the slope, the
unconformities pinch together and cannot be distinguished from one
another. Further, taking account of the widespread correlation of
unconformities (on widely separated continental margins) and the Vail
coastal onlap curve (Vail et al., 1977; Vail and Hardenbol, 1979), the DSDP
holes along line 25 offer much support for the idea that sea-level changes
were a major factor during the late Mesozoic and Cenozoic in controlling
sedimentation and erosion for margins bordering the North Atlantic.
The establishment through drilling (especially at DSDP Site 612; see
Figure 1), of a close relation between unconformities as seen on seismic
reflection profiles and those identified in boreholes thereby validates the

Vail depositional model as a useful predictive tool in this geological frontier.


These sea-level changes are in addition to gradual changes in paleoclimate,
seawater temperature, and global ice volumes that affected the trend of
global sea level.
The drilling further documented the present shelf-slope-rise transition as
a relatively recent (Pleistocene) feature. The data also indicate that during
the early Cenozoic, the transition was much broader and included a
backshelf and foreshelf (Poag, 1986); major shifts in the depositional
regime took place during the Paleocene, Eocene, Miocene, and Pleistocene.
The slope area built up as a result of the complex interplay of deposition of
shelf-derived sediments and erosion by geostrophic currents, and by
downslope gravity-induced sedimentary processes. On the rise fans,
mounded drift deposits and hemipelagic(?) blankets of basin fill developed
during this time.

REFERENCES CITED
Grow, J.A., D.R. Hutchinson, K.D. Klitgord, W.P. Dillon, and J.S. Schlee, 1983,
Representative multichannel seismic profiles over the U.S. Atlantic
margin, in A.W. Bally, ed., Seismic expression of structural styles: AAPG
Studies in Geology Series 15, v. 2, p. 2.2.3-1-19.
Poag, C.W., 1987, The New Jersey transect: stratigraphic framework and
depositional history of a sediment-rich passive margin, in C.W. Poag, A.B.
Watts, et al., Initial reports of the Deep Sea Drilling Project: Washington,
D.C., U.S. Government Printing office, v. 95, p. 763-817.
Schlee, J.S., and K. Hinz, 1987, Seismic stratigraphy and facies of continental
slope and rise seaward of Baltimore Canyon trough: AAPG Bulletin, v. 71,
p. 1046-1067.
Schlee, J.S., C.W. Poag, and K. Hinz, 1985, Seismic stratigraphy of the continental
slope and rise seaward of Georges Bank, in C.W. Poag, ed.,Geologic evolution of the
United States Atlantic margin:New York,Nostrand Co., p. 265-292.
Vail, P.R., R.M. Mitchum, R.G. Todd, J.M. Widmier, S. Thompson, J.B. Sangree,
J.N. Bubb, and W.G. Hatlelid, 1977, Seismic stratigraphy and global
changes in sea level, in C.E. Payton, ed., Seismic stratigraphy
applications to hydrocarbon exploration: AAPG Memoir 26, p. 49-212.
Vail, P.R., and J. Hardenbol, 1979, Sea level changes during the Tertiary: Oceanus,
v. 22, n. 3, p. 71-79.

Table 1. Inferred and documented (through drilling) reflector ages and stratigraphic
relations.

Reflector

Approximate Age

1
2A
2B
2C

Base of the Quaternary


Base of the Pliocene
Middle Miocene-Upper Miocene
Lower Miocene-Middle Miocene

2D
2X
2Y
3

Oligocene-Lower Miocene
Middle Eocene-Middle Miocene
Paleocene-Maestrichtian
Top of Cretaceous

4
5
6
7
8

Top of Jurassic
Middle-Upper Jurassic(?)
Lower-Middle Jurassic(?)
Lower-Middle Jurassic(?)
Lower-Middle Jurassic(?) top of oceanic basement

Line 25 across New Jersey shelf, slope and upper rise

48

BALTIMORE CANYON TROUGH, MID-ATLANTIC OCS:


SEISMIC STRATIGRAPHY OF SHELL/AMOCO/SUN WELLS

R.N. ERLICH, K.P. MAHER, G.A. HUMMEL, D.G. BENSON,


G.J. KASTRITIS, H.D. LINDER, R.S. HOAR
and D.H. NEELEY
Amoco Production Company
New Orleans, Louisiana

Seismic line A-A transects the Civet and Rhino well locations and
shows the structural and stratigraphic settings of each (Figures 4, 5, 6, 7,
and 8). Drilling problems made sampling most of the Tertiary-Quaternary
section in the Civet, Rhino, and Hyena wells impossible. However, the
early-middle Oligocene Au disconformity of Tucholke and Mountain (1979)
is clearly visible seismically and generally correlates with some reworked
Eocene sections samples in the Civet well (Figure 3). The late Miocene
erosional event described by Van Hinte et al. (1985a) at DSDP Sites 604 and
605 is not easily defined on seismic lines A-A or B-B (Figures 9, 10, 11,
and 12), although it is seen on lines from the southern Baltimore Canyon
Trough.

INTRODUCTION
The structural and stratigraphic history of the Baltimore Canyon Trough
is recently the topic of much study (Schlee, 1981; Libby-French, 1984;
Edson, 1985; Gamboa et al., 1985; Poag, 1985; Van Hinte et al., 1985a and
b). Many have proposed the existence of a Lower Cretaceous-Upper
Jurassic reef complex underlying the present-day continental slope, but little
direct supporting evidence was available until now.
Shell, Amoco, and Sun jointly drilled three exploratory wells in the
Baltimore Canyon Trough to test prospective Lower Cretaceous and Upper
Jurassic reef and back-reef structures (Figures 1 and 2). Drilling began in
August, 1983, on Wilmington Canyon Block 587, with the Civet well,
followed by the Rhino well on Block 586. Total depth for the Civet well
was 14,500 ft (4394 m), and total depth for the Rhino well was 16,000 ft
(4848 m). Evaluation of carbonate prospects ended in July, 1984, with
completion of the Hyena well on Block 372. Total depth for the Hyena
well was 11,630 ft (3524 m). The wells encountered no commercial
hydrocarbon shows.

SEISMIC STRATIGRAPHY
General Geology
Cuttings, conventional cores, and seismic data were used to develop a
seismic stratigraphic and facies framework for the Civet, Rhino, and Hyena
wells and adjacent areas. Age was determined using palynology and
foraminifer biostratigraphy. These ages were compared to data from other
industry wells in the Baltimore Canyon Trough and were tied (seismically)
back to existing well control (Figure 3). These seismic/well ties yielded
regionally correlative and continuous mappable units.

R.N. Erlich et al.

Figure 1. Index map to study area: Lower Cretaceous-Jurassic reef trend, eastern
North America.

Upper Cretaceous sediments encountered in the Civet, Rhino, and


Hyena wells are mostly pelagic shales and globigerinid oozes (Figure 3).
Both the Civet and Rhino wells encountered a few thin quartzose sands in
the lower-middle Cenomanian. Unconformities of Turonian-Coniacian,
Coniacian-Santonian, and Santonian-Campanian age were noted in the
COST B-3 well by Poag (1985), but were not noted in the Civet, Rhino, or
Hyena wells. Albian-Aptian and lower Neocomian unconformities were
found in the Civet well using palynology, but the Albian-Aptian
unconformity was the only unconformity noted in the Rhino and Hyena
wells. An expected unconformity at the top of the Jurassic section was not
present in the Civet or Rhino wells.
Albian-Aptian rocks in the Rhino well consist of quartz sandstone that
grades upward into oolitic and lagoonal limestone (Figures 5, 7, and 13).
This Albian-Aptian carbonate section extends at least 30 mi (36 km) north
and 10 mi (12 km) south of the Rhino well, and contains numerous oolite
shoals and patch reefs (Figures 10 and 12). This shoal- and patch-reef trend
may overlie the last phase of an Aptian-Hauterivian deltaic sequence, which
apparently built a large deep-water submarine fan near DSDP Site 603 (Van
Hinte et al., 1985b). This deltaic sequence also might have severely reduced
or eliminated pre-AlbianAptian carbonate deposition in most of the
northern Baltimore Canyon Trough.
The Valanginian-Berriasian sections of the Civet and Rhino wells
consist of impermeable, shallow-water lagoonal limestones with a few thin
shale interbeds (Figures 3 and 13). The Hyena well also penetrated about
250 ft (76 m) of continuous lower Aptian-Hauterivian lagoonal limestones
and about 450 ft (136 m) of boundstones and grainstones, interpreted as the
upper part of a Tithonian-Neocomian shelf-edge reef complex. The well
reached total depth in this facies without encountering Jurassic rocks
(Figures 10, 12, and 13).
Jurassic rocks in the Civet well range from dense lagoonal limestones
and porous, skeletal packstones and grainstones in the upper Tithonian, to
dense, stylolitic foreslope limestones in the lower Tithonian (Figures 8 and
13). The Civet well reached total depth in these rocks.
The Tithonian section of the Rhino well closely resembled that of the
Civet well, but contained fewer coarse-grained limestones with more
abundant quartz sandstone and shale interbeds, especially near the

Figure 2. Locations of wells in the Baltimore Canyon Trough; index to seismic lines.

Baltimore Canyon trough, seismic stratigraphy

51

Kimmeridgian-Tithonian boundary. Seismically, this boundary appears as a


flat disconformity over prograding clinoforms (Figures 5, 7, 8, and 13). No
paleontologic evidence supporting a depositional hiatus was found at the
boundary, so the exact nature of this reflector remains unclear.
Kimmeridgian rocks in the Rhino well range from dense, lagoonal
limestones with thin sandstone and shale interbeds at the top of the
Kimmeridgian, to dense, oolitic-skeletal packstones and grainstones at total
depth (Figures 8 and 13). Palynological data show that no rocks older than
upper Kimmeridgian were encountered in the Civet, Rhino, or Hyena wells.
Facies Models
Carbonate deposition in the Baltimore Canyon Trough in the upper
Kimmeridgian-lower Tithonian probably occurred on a low-to-moderateenergy accretionary margin; upper Tithonian-lower Neocomian carbonates
were probably deposited on a moderate-energy bypass margin (Figures 8
and 14). These platform margin types appear to be similar to those
described by McIlreath and James (1979) and Read (1985). Platform
margin facies evolved from ooid sand and skeletal sand shoals in the
Kimmeridgian to discontinuous reefs and shoals in the upper Tithonianlower Neocomian. Upper Tithonian-lower Neocomian facies patterns show
that reef frameworks were dominated by a stromatoporoid-sponge
assemblage with minor corals (Figure 15). This assemblage is similar to
some upper Jurassic reefs described by Turnsek et al. (1981) in Yugoslavia.
Lower Tithonian foreslope facies contained common Tubiphytes-algal
mounds similar to those found in the subsurface off Nova Scotia (Eliuk,
1978).
Seismic and well data show that platform margin carbonate deposition
in the Baltimore Canyon Trough did not occur in the same geographic
position continuously from Oxfordian through Valanginian, as Poag
proposed (1985). His model suggests that an Oxfordian-Valanginian shelfedge reef with at least 5000 to 8000 ft (1500 to 2000 m) of relief directly
underlies the Tithonian-lower Neocomian platform margin at about 16,500
ft (5000 m) subsea. Data presented in this study show that platform margin
deposition at the Baltimore Canyon Trough is probably no older than lower
Tithonian, and that relief on the margin, as determined from seismic data,
was probably never greater than about 2000 ft (606 m). Most of this relief
probably occurred during the upper Tithonian-lower Neocomian (Figures 7,
8, and 12).
Changes in the location of platform margin deposition and facies
suggest that Jurassic-lower Neocomian platform margin facies changes in
the Baltimore Canyon Trough might have been caused by local
modifications of eustatic sea level fluctuations (Vail et al., 1984). Schlager
and Ginsburg (1981) also attributed similar facies changes in Bahamian,
Tertiary-Quaternary limestones to rapid sea level changes. Therefore, rapid

R.N. Erlich et al.

shelf margin progradation between the upper Kimmeridgian and lower


Tithonian was probably caused by a sudden, but short-lived, sea level fall
(Figure 16). Rapid vertical upbuilding of the Tithonian-lower Neocomian
shelf margin probably was caused by a rapid transgression that persisted
into the lower Neocomian (Berriasian). A major regressive event occurred
near the end of the lower Neocomian and allowed large delta systems to
build out and bury most of the upper Tithonian-lower Neocomian carbonate
platform (Figure 13). A gradual transgression followed and lasted until the
middle Aptian, reworking delta front deposits along depositional strike with
the Civet, Rhino, and Hyena wells. A rapid regression followed this
transgressive event. This regression, combined with a lack of deltaic
deposition, reestablished carbonate deposition over most of the Baltimore
Canyon Trough. Prograding oolite shoals and patch reefs were subsequently
drowned during a transgression that began in the early Albian and lasted
well into the Cenomanian (Vail et al., 1984).
Detailed seismic stratigraphic, facies, and structural modeling suggest
that local variations in the rates of sediment supply and basin subsidence
have modified eustatic sea level changes in the Baltimore Canyon Trough.
This has produced a complex series of local onlapping and offlapping units,
each with a characteristic seismic response. A more detailed comparison
between these shelf and shelf-margin units and recent DSDP well data (Van
Hinte et al., 1985a, 1985b) will yield a precise depositional history for this
part of the mid-Atlantic outer continental shelf.

SEISMIC ACQUISITION AND PROCESSING


One hundred twenty channels of 9-sec, 60-fold seismic data were
recorded in late 1983 using a 2 msec sampling rate. Group and source
intervals of 82 ft (25 m) produced a CDP interval of 41 ft (12 m). Near
offset to the spread was 820 ft (248 m) and far offset was 10,580 ft (3206 m).
The data were demultiplexed and resampled to 4 msec and were
amplitude edited and muted. These procedures were followed by spherical
divergence correction. The field data were then sorted to 60-fold and an
inside mute was used to remove some of the water-bottom multiple energy.
Normal moveout correction was followed by a 60-fold stack. The stacked
data were then filtered with a time variant filter that followed the water
bottom. Long window automatic gain control (AGC) scaling was used to
prepare the data for migration, which was then done using a Kirchoff
migration algorithm. Finally, the data were rescaled with a 500 msec AGC
and plotted using standard Amoco plotting procedures.

ACKNOWLEDGMENTS
The authors wish to thank Rudy Lippert and Chick Voorhies for their
valuable contributions during the early days of our exploration program. We
are indebted to Clarence Jamison for his tireless efforts in helping prepare
the illustrations, and to Amoco Production Company for providing
permission to publish this report.

REFERENCES CITED
Edson, G.M., 1985, The mid-Atlantic paleoshelf edge-carbonate buildup or
reef? (abs.): AAPG Bulletin, v. 69, p. 1436.
Eliuk, L.S., 1978, The Abenaki Formation, Nova Scotia shelf, Canada-a
depositional and diagenetic model for a Mesozoic carbonate platform:
Bulletin Canadian Petroleum Geology, v. 26, p. 424-514.
Gamboa, L.A., M. Truchan, and P.L. Stoffa, 1985, Middle and Upper Jurassic
depositional environments at outer shelf and slope of Baltimore Canyon
Trough: AAPG Bulletin, v. 69, p. 610-621.
Libby-French, J., 1984, Stratigraphy, framework, and petroleum potential of
northeastern Baltimore Canyon Trough, mid-Atlantic Outer Continental
Shelf: AAPG Bulletin, v. 68, p. 50-73.
McIlreath, I.A., and N.P. James, 1979, Facies models, 12: carbonate slopes, in R.G.
Walker, ed., Facies models: Geoscience Canada Reprint Series 1, p. 133-143.
Poag, C.W., 1985, Depositional history and stratigraphic reference section for
central Baltimore Canyon Trough, in C.W. Poag, ed.,Geologic evolution of the
United States:Atlantic margin:New York,Van Nostrand Reinhold Co., p. 217-264.
Read, J.F., 1985, Carbonate platform facies models: AAPG Bulletin, v. 69, p. 1-21.
Schlee, J.S., 1981, Seismic stratigraphy of Baltimore Canyon Trough: AAPG
Bulletin, v. 65, p. 26-53.
Schlee, J.S., 1981, Seismic stratigraphy of Baltimore Canyon Trough : AAPG
Bulletin, v. 65, p. 26-53
Schlager, W., and R.N. Ginsburg, 1981, Bahama carbonate platforms-the deep
and the past: Marine Geology, v. 44, p. 1-24.
Tucholke, B.E., and G.S. Mountain, 1979, Seismic stratigraphy, lithostratigraphy,
and paleosedimentation patterns in the North American basin, in M. Talwani, W.
Hay, and B.F. Ryan,eds.,Deep drilling results in the Atlantic Ocean: American
Geophysical Union Maurice Ewing Series, Symposium Proceedings 3, p. 58-86.
Turnsek, D., S. Buser, and B. Ogorelec, 1981, An upper Jurassic reef complex
from Slovenia, Yugoslavia, in D.F. Toomey, ed., European fossil reef
models: SEPM Special Publication 30, p. 361-369.
Vail, P.R., J. Hardenhol, and R.G. Todd, 1984, Jurassic unconformities,
chronostratigraphy, and sea level changes from seismic stratigraphy and
biostratigraphy, in J.S. Schlee, ed., Interregional unconformities and
hydrocarbon accumulation: AAPG Memoir 36, p. 219-144.
Van Hinte, J.E., et al, 1985a, Deep sea drilling on the upper continental rise off
New Jersey; DSDP Sites 604 and 605: Geology, v. 13, p. 397-400.
Van Hinte, J.E., et al, 1985b, DSDP Site 603: First deep (E1000 m) penetration of
the continental rise along the passive margin of eastern North America:
Geology; v. 13, p. 392-396.

Baltimore Canyon trough, seismic stratigraphy

52

TERTIARY DEPOSITIONAL SEQUENCES,


OFFSHORE NEW JERSEY AND ALABAMA

S.M. GREENLEE
Exxon Production Research Company
Houston, Texas
INTRODUCTION
Tertiary strata beneath the continental shelf offshore New Jersey and
Alabama are ideally suited for seismic stratigraphic analysis. Both areas are
drilled and studied paleontologically, enabling accurate dating of
depositional sequences recognized on seismic profiles. Facies analysis from
paleontological studies, well logs, and sample cuttings from these offshore
wells help constrain and calibrate seismic facies interpretations. Subsidence
was relatively slow and continuous, and neither area has been severely
deformed by basement-involved or detached faulting during the Tertiary.
Finally, excellent seismic data are available from both areas.
The offshore New Jersey study area was the locus of deep-water
carbonate-rich sedimentation during the Paleogene until the late Oligocene.
Prograding siliciclastics characterize deposition in the area during the
Neogene, when deltaic wedges built seaward to the present-day shelf edge.
The off-shore Alabama study area also was in a distal position until the late
Oligocene. Following a period of late Oligocene to early Miocene reef
growth, this area was buried by thick, prograding siliciclastic wedges.
Because of the extensive data base and the generally similar
paleogeographic setting, depositional systems, and structural history of
these two areas, they provide an excellent opportunity to compare the
character, timing, and extent of depositional sequences found on two
different parts of the North American continental margin.

REGIONAL GEOLOGIC SETTING


Previous studies document the geological evolution of the offshore New
Jersey area (Grow, 1980; Schlee, 1981; and Poag, 1985). Seismic profiles
from this area appear in the previous AAPG atlas titled Seismic Expression
of Structural Styles (Bally, 1983) through contributions by Lippert, Grow et
al., Crutcher, and Morgan and Dowdall. A more detailed discussion of the

S.M. Greenlee

Figure 1. Location of seismic sections illustrated, offshore wells, and positions of


Neogene depositional shelf edges, offshore New Jersey.

Tertiary depositional sequences recognized on this grid and their


relationship to eustatic cycles is found in Greenlee et al. (1987) and Moore et al.
(1987). Location of wells and lines discussed in this chapter are shown in Figure 1.
Much less information exists in the geologic literature concerning the
Tertiary of the offshore Alabama area. A seismic section from the seaward
part of the area is shown in the structural styles atlas (paper by Fluker).
Studies of Tertiary strata are (presently) restricted to coastal plain sections
(see Murray, 1961; Toulmin, 1977; Mancini, 1981; Raymond, 1985).
Tertiary strata prograde basinward over a paleotopography controlled by
Early Cretaceous carbonate deposition. The position of the buried Early
Cretaceous bank margin is shown in Figure 2. This separates relatively flatlying platform-interior strata from steeply dipping slope strata. Tertiary
progradation reached this relict shelf edge in middle Miocene time (Figure
2) when sediment derived from the north and east filled in most of the
preexisting shelf-to-basin physiography. Late Miocene and younger
siliciclastic progradation built the edge of the continental shelf to its present position.
Two geohistory plots (see Van Hinte, 1978) illustrate the subsidence
history of key well locations in the study area (Figure 3). The first, from the
COST B-2 well off New Jersey (Figure 1), is taken from Greenlee et al.
(1987). It shows two episodes of uplift followed by exponentially
decreasing rates of thermo-tectonic subsidence. These uplift episodes are
interpreted to represent the initial rifting of North America from Africa and
an Early Cretaceous period of volcanism associated with emplacement of
the Great Stone Dome. Slow, continuous subsidence characterizes the
Tertiary at this location.
The second geohistory diagram is from the Shell Main Pass 154 well
(see Figure 2 for location). A Cretaceous uplift event also is recognized
here; however, a marked increase in the rate of thermo-tectonic subsidence
occurs in the middle Miocene. This increase in subsidence is interpreted to
be an effect of flexural downwarping of the lithosphere due to extremely
rapid middle Miocene sedimentation seaward of the Lower Cretaceous bank
margin. Despite this increase in subsidence there is no evidence for a
precursor uplift event. Because both areas experienced continuous
subsidence throughout the Tertiary, downward shifts in coastal onlap below
the previous depositional shelf margin are interpreted as a sedimentary
response to a drop in eustatic sea level.

TERTIARY DEPOSITIONAL SEQUENCES


Depositional sequences are recognized on seismic sections by erosion
and basinward shifts in coastal onlap (see Vail, this title, volume 1).
Although no coastal onlap can be identified in either area until the
Oligocene, sequence boundaries are recognized on the basis of erosional
truncation of underlying reflections. Downlap surfaces also are noted on the
seismic sections by dashed lines. Biostratigraphic analysis of available wells
is then used to date the depositional sequences. Offshore New Jersey wells
with biostratigraphic analysis are shown in Figure 1; these wells were
studied by Poag (1980, 1985) and Exxon paleontologists. Offshore Alabama
wells noted in Figure 2 also were studied by Exxon paleontologists.
Age-dated depositional sequences are then correlated with eustatic
cycles represented on the eustatic cycle chart (Haq et al. (1987) (Figure 4).
The numbers used to denote the sequence boundaries on the seismic section
result from this correlation. During the Paleogene when the locus of coastal
deposition was distal with respect to the study areas, only second-order or
supersequences may be recognized on the seismic sections (e.g. TA1,
TA2, etc.). Nearer to the depocenter, third-order sequences are recognized.
The Cretaceous-Tertiary boundary is an erosional surface on both
margins, with more severe erosion occurring offshore Alabama. It is
overlain by a series of downlapping clinoforms. The downlap surface at the
Cretaceous-Tertiary boundary is interpreted to indicate a rapid rise in sea
level in the latest Maastrichtian and earliest Paleocene. In the offshore New
Jersey area, the entire lower Paleocene section is absent because of
condensed sedimentation and mid-Paleocene erosion. This condensed
section is noted at the Cretaceous-Tertiary boundary on Figure 4.
Sequence boundaries characterized by planar erosional surfaces with
varying degrees of deep marine onlap occur near the top of the lower and
middle Eocene in both areas. These are correlated with second-order
sequence boundaries at 49.5 m.y. and 39.5 m.y. A sequence boundary of
upper Paleocene age (58.5 m.y.) appears in the offshore Alabama grid but
has merged with the Cretaceous-Tertiary boundary in the offshore New
Jersey area. Both areas exhibit onlap of upper Eocene sediments against the
39.5 m.y. sequence boundary.
A major downward shift in coastal onlap in mid-Oligocene time
established coastal sedimentation in both areas. This is followed by

Tertiary depositional sequences, offshore New Jersey and Alabama

67

progressive landward onlap of upper Oligocene and lower Miocene


(Aquitanian) deposits. In the offshore Alabama area, the lowermost
onlapping deposits are sandstones overlain by an uppermost Oligocene and
lower Miocene carbonate unit. A coral reef penetrated at Viosca Knoll 30 is
shown in Figure 8. These carbonates extend landward beyond the seismic
grid. Reef growth was terminated by a major downward shift in onlap at

Figure 2. Location of seismic sections illustrated, offshore wells used in study,


and positions of Lower Cretaceous and Neogene shelf edges, offshore Alabama.

S.M. Greenlee

21 m.y. Coeval sequences offshore New Jersey are composed of prograding


siliciclastics that show a similar coastal onlap pattern. Upper Oligocene
sediments here also are primarily restricted to basinal areas and, together
with lower Miocene deposits, onlap landward until they extend beyond the
study grid by late Aquitanian time.
A downlap surface at 18.5 m.y. is recognized in both areas. Uppermost

lower Miocene and lower middle Miocene sequences extend landward


beyond both seismic grids. These sequences are more clearly expressed in
the offshore New Jersey area, which was characterized by a greater amount
of sediment influx during this time. The 15.5 m.y. sequence boundary is a
significant erosional surface in both areas. In offshore New Jersey (Figures
5 and 6), this planar erosional surface truncates progressively older strata in
a basinward direction while offshore Alabama shows significant erosion
beneath this surface above the relict Lower Cretaceous shelf margin (Figure 8).
A significant departure in coastal onlap among the two areas occurs in
the middle Miocene following the 15.5 m.y. sequence boundary. Whereas
both show strongly progradational middle Miocene depositional sequences
over a downlap surface at 15 m.y., those in the offshore Alabama area are
more extensive areally and clearly extend well updip of the seismic grid.
All middle Miocene sequences are well developed on both seismic grids
and are composed of a basinally restricted wedge that often is
progradational in character, and an areally extensive regressive wedge that
progrades over the top of the basinally restricted wedge. These two
components are interpreted to represent the lowstand wedge systems tract
(deposited when sea level was at or below the preexisting depositional shelf
edge), and the highstand systems tract (deposited during the subsequent
highstand in sea level; see Vail, this title, volume 1). In some cases, a thin
interval of sediment that extends landward of the previous shelf edge may
be resolved; it is overlain by the downlapping clinoforms of the succeeding
highstand systems tract (see Figure 8). This is interpreted as the
transgressive systems tract.
A sequence boundary near the top of the middle Miocene is recognized
on both grids as the most pronounced downward shift in coastal onlap in the
Neogene. In New Jersey, late Miocene deposits onlap to the west and north
against the uppermost middle Miocene sequence boundary. Miller et al.
(1987) documented canyon cutting associated with this erosional event in
New Jersey. The lower upper Miocene sequence is basinally restricted and
is best developed to the south (Figure 1). Younger late Miocene sequences
onlap progressively, further landward. The offshore Alabama area (Figures
7 and 8) was dissected by numerous submarine canyons during this time,
one of which is crossed in an axial position by the seismic line in Figure 7.
A mounded package of sediment covers this canyon floor and is interpreted
as debris from headward erosion of the canyon walls. Late Miocene
sequences are thickest seaward of the previous depositional shelf edge, and
have thin updip extensions.
Early Pliocene sequences are areally extensive in both study areas and
extend far updip. Sequence cycles are more frequent offshore Alabama than
are noted on the global cycle chart (Haq et al. (1987; also Figure 4). Four
depositional sequences of this age are recognized in Alabama. In New
Jersey, the Pliocene sedimentary section is thinner and detailed
interpretation is hindered by water-bottom multiple interference.

CONCLUSIONS
Seismic stratigraphers use coastal onlap patterns to compare the shifts in
areal extent of depositional sequences with time on widely separated
continental margins (see Vail et al. 1977; Vail, this title, volume 1).
Examination of coastal onlap patterns from offshore New Jersey and
Alabama (Figure 3) shows several key similarities.
First, a mid-Oligocene downward shift in onlap followed by progressive
landward onlap through the late Oligocene, and early and middle Miocene,
occurs on both margins. This onlapping series of depositional sequences
shows cyclic downward shifts in onlap near the sediment depocenters,
which we interpret to represent correlative third-order sequence boundaries.
A more basinward onlap position for the 13.8 to 15.5 m.y. sequence in the
offshore New Jersey area may be a function of greater subsidence occurring
at this time in offshore Alabama (Figure 4), erosion of the sequence in
updip areas of the New Jersey study area, or an uplift event not recognized
in the New Jersey geologic history.
Another major basinward shift in onlap exists near the top of the middle
Miocene, which results in a basinal restriction of lower upper Miocene
sequences. A thin, landward extension of the offshore Alabama lower upper
Miocene wedges is probably due to the greater subsidence of this area.
Lower Pliocene deposits are widespread in both areas. Other notable
similarities include the timing and character of second-order Paleogene
sequence boundaries and the major downlap surface at the base of the
Tertiary.
Recognition of the common attributes of the timing, areal distribution,
and character of Tertiary depositional sequences in these two areas
characterize the observations that led Vail and others (Vail et al., 1977) to
propose shifts in eustatic sea level as a control on global sedimentary
patterns. Although each basin is characterized by unique tectonism and
depositional systems, an analysis of the depositional sequences relative to
global cycles noted on the eustatic cycle chart provides an additional tool
for pre-drill age, and lithofacies prediction.

ACKNOWLEDGMENTS
Exxon paleontologic reports form a key data base for this study, and the
contributions of Exxon U.S.A. paleontologists, especially Marilyn Crane,
are gratefully acknowledged. Exxon Production Research Co. geologists, in
particular P.R. Vail, T.S. Loutit, M.G. Fitzgerald, and E.W. Schroeder made
contributions to understanding the stratigraphy of these two areas. I thank
Exxon Company, U.S.A., and Exxon Production Research Co. for
permission to publish this paper.

Tertiary depositional sequences, offshore New Jersey and Alabama

68

REFERENCES CITED

Figure 3. Geohistory diagrams of (a) the COST B-2 well (offshore New Jersey) and (b) the Shell Main Pass 154
well (offshore Alabama). The lower curve represents total subsidence of the basement through time; the upper
curve represents total subsidence corrected for sediment load and compaction effects. The uppermost curve
represents a long-term sea level curve (Haq et al., 1987) used as a datum, and the shaded area represents
paleowater depth interpretation.

S.M. Greenlee

Bally, A.W., 1983, Seismic expression of structural stylesa picture and work
atlas: AAPG Studies in Geology Series 15, 3 volumes.
Greenlee, S.M., F.W. Schroeder, and P.R. Vail, 1987, Seismic stratigraphic and
geohistory analysis of Tertiary strata from the continental shelf off New
JerseyCalculation of eustatic fluctuations from stratigraphic data, in
R.E. Sheridan, ed., The geology of the Atlantic margin: Geological Society
of America, Decade of North American Geology Series (in press).
Grow, J.A., 1980, Deep structure and evolution of the Baltimore Canyon Trough in
the vicinity of the COST No. B-3 well, in P.A. Scholle, ed., Geological
studies of the COST No. B-3 well, United States and mid-Atlantic
continental slope area: U.S. Geological Survey Circular 833, p. 117-132.
Haq, B.U., J. Hardenbol, and P.R. Vail, 1987, Chronology of fluctuating sea levels
since the Triassic: Science, v. 235, p. 1156-1167.
Mancini, E.A., 1981, Lithostratigraphy and biostratigraphy of Paleocene
subsurface strata in southwest Alabama: Gulf Coast Association of
Geological Societies Transactions, v. 31, p. 359-367.
Miller, K.G., A.J. Melillo, G.S. Mountains, and J.A. Farre, 1987, Middle/late
Miocene canyon cutting on the New Jersey continental slope
biostratigraphic and seismic stratigraphic evidence:preprint submitted to Geology.
Moore, T.C., T.S. Loutit, and S.M. Greenlee, 1987, Estimating short-term changes
in eustatic sea level: in preparation.
Murray, G.E., 1961, Geology of the Atlantic and Gulf Coast province of North
America: New York, Harper and Brothers, 692 p.
Poag, C.W., 1980, Foraminiferal stratigraphy, paleoenvironments, and depositional
cycles in the outer Baltimore Canyon Trough, in P.A. Scholle, ed.,
Geological studies of the COST No. B-3 well, United States mid-Atlantic
continental slope area: U.S. Geological Survey Circular 833, p. 44-65.
Poag, C.W., 1985, Depositional history and stratigraphy reference section for
central Baltimore Canyon Trough, in C.W. Poag, ed., Geologic evolution
of the United States Atlantic margin:New York,Van Nostrand Reinhold, p. 217-264.
Raymond, D.E., 1985, Depositional sequences in the Pensacola Clay (Miocene) of
south-west Alabama: Geological Survey Alabama Bulletin 114, 87 p.
Schlee, J.S., 1981, Seismic stratigraphy of the Baltimore Canyon Trough: AAPG
Bulletin, v. 65, p. 2653.
Toulmin, L.D., 1977, Stratigraphic distribution of Paleocene and Eocene fossils in
the eastern Gulf Coast region: Alabama Geological Survey Monograph 13,
v. 1, 602 p.
Vail, P.R., R.M. Mitchum, Jr., and S. Thompson, III, 1977, Seismic stratigraphy
and global changes of sea level, Part 4; Global cycles of relative changes
of sea level, in C.E. Payton, ed., Seismic stratigraphyapplications to
hydrocarbon exploration: AAPG Memoir 26, p. 83-97.
Vail, P.R., J. Hardenbol, and R.G. Todd, 1984, Jurassic unconformities,
chronostratigraphy and sea-level changes from seismic stratigraphy and
biostratigraphy, in J.S. Schlee, ed., Interregional unconformities and
hydrocarbon accumulation: AAPG Memoir 36, p. 129-144.
Van Hinte, J.E., 1978, Geohistory analysisapplication of micro-paleontology in
exploration geology: AAPG Bulletin, v. 62, p. 201-222.

Tertiary depositional sequences, offshore New Jersey and Alabama

69

U.S. GEOLOGICAL SURVEY LINE 19


ACROSS THE GEORGES BANK BASIN

K.D. KLITGORD and J.S. SCHLEE


U.S. Geological Survey
Woods Hole, Massachusetts
and
J.A. GROW
U.S. Geological Survey Denver, Colorado
INTRODUCTION
Georges Bank is a shallow part of the Atlantic continental shelf
southeast of New England (Emery and Uchupi, 1972, 1984). This bank,
however, is merely the upper surface of several sedimentary basins
overlying a block-faulted basement of igneous and metamorphic crystalline
rock. Sedimentary rock forms a seaward-thickening cover that has
accumulated in one main depocenter and several ancillary depressions,
adjacent to shallow basement platforms of paleozoic and older crystalline
rock. Georges Bank basin contains a thickness of sedimentary rock greater
than 10 km, whereas the basement platforms that flank the basin are areas
of thin sediment accumulation (less than 5 km).

STRUCTURAL FRAMEWORK
The primary structural elements of the Georges Bank basin region are
typical of passive continental margins: shallow basement platform, deep
marginal sedimentary basin, and deep ocean basins (Figure 1). The tectonic
evolution of a passive margin creates distinctive crustal structures, basement
structures, and sediment-distribution patterns. Passive continental margins
form as the result of a continent breaking apart (continental rifting) and then
moving apart (continental drift) to create a new ocean basin (sea-floor
spreading). During the rifting phase of margin development, crustal
stretching, thinning, and block faulting take place along the rift zone as the
two parts of a plate slowly move apart. Extensional tectonic activity in the
rift terminates when the extensional plate boundary moves seaward and the
seafloor-spreading process creates new oceanic crust at a mid-ocean ridge.

K.D. Klitgord, J.S. Schlee and J.A. Grow

A new, deep-ocean basin continues to widen during the sea-floor-spreading


phase, and margin development is then controlled by thermal and sedimentloading tectonic processes interacting with oceanographic processes.
In the Georges Bank region, crystalline basement includes Paleozoic
and older rocks on the platforms, block-faulted pre-Mesozoic rocks mixed
with Triassic and Jurassic igneous rocks beneath the marginal basin, and
Jurassic and younger igneous rocks seaward of the marginal basin.
Basement samples are few, and most information about basement is inferred
from seismic, magnetic, and gravity data.
Seismic-reflection profiles like line 19 show great variability in depth
and character of acoustic and crystalline basement. Crystalline basement
can be divided into the following zones: (1) a low-relief zone of Paleozoic
or older metamorphic and igneous rocks forming continental basement
(upper surface of continental crust) on the shallow stable platforms; (2) a
basement hinge zone (blockfaulted zone of Paleozoic or older metamorphic
and igneous rocks that deepens seaward in distinct steps; see line 19); (3)
the marginal sedimentary basin where crystalline basement is masked by
thick sedimentary units that contain prograding carbonate deposits at the
seaward edge; and (4) a zone in the deep ocean basin of igneous rocks
forming oceanic basement (upper surface of oceanic crust) produced by seafloor spreading.
The block-faulted zone southeast of the Gulf of Maine platform (Figure
1) deepens in a series of steps into a broad crustal depression (Yarmouth
sag) adjacent to Yarmouth arch. Nestled between each step is a small
subbasin formed by the downdropped block (graben); over these subbasins,
the postrift unconformity separates steeply dipping older sedimentary
reflectors from more conformable, flat-lying sedimentary units (see Line
19).
Yarmouth arch is separated from the Gulf of Maine platform blockfaulted zone by the Yarmouth sag. The Yarmouth arch and sag form the
basement transition between the LeHave platform and the main Georges
Bank basin, and at their eastern end they appear to form part of the Gulf of
Maine platform block-faulted zone. They deepen to the southwest until
CDP Line 19, where the arch is inferred to be only a small basement high
that has a buildup of carbonate rock on top (Poag, 1982), and the sag is at
the seaward edge of the Gulf of Maine platform block-faulted zone.

Crustal Structure
The three primary crustal types in the Georges Bank region are
continental, transitional (or rift stage), and oceanic crust; they coincide with
the primary structural elementsGulf of Maine platform, the block-faulted
zone and Georges Bank basin, and the deep ocean basin. Continental crust
is thickest and underlies the basement platforms. Oceanic crust is the
thinnest and underlies oceanic basement in the deep-ocean basin. Crust of
the deep marginal deep sedimentary basin is of intermediate thickness, and
its nature is the most speculative. Estimates of crustal thickness and type
come from seismic-refraction profiles (Sheridan et al., 1979), magnetic
studies (Klitgord and Behrendt, 1979; Klitgord et al., 1982), gravity models
(Grow et al., 1979; Swift et al., 1987), and thermal subsidence models
(Sawyer et al., 1982, 1983). A major magnetic lineation, the East Coast
Magnetic Anomaly (ECMA), marks the landward edge of oceanic crust and
is located over the upper continental rise adjacent to Georges Bank. Gravity
model studies indicate that the continental crust under the platform (about
35 km thick) thins rapidly at the basement hinge zone and reaches oceanic
crustal thickness (about 5 km) at the ECMA (Figure 17). In broader sections
of the marginal basin, gravity modeling suggests that the crustal thickness
thins rapidly at the hinge zone, becomes more uniform beneath the marginal
basin, and then thins again at the ECMA.
Structures within the crust are most readily seen on our seismic profiles
across the platform and basement hinge zone. Crust-cutting faults can be
traced downward from the border faults of the half-graben structures at the
hinge zone of Georges Bank basin and on the edges of the Nantucket and
Atlantis basins (Figure 1). These features are interpreted as normal faults
that formed in response to extensional tectonic forces during the rifting
phase of margin development; some may be reactivated Paleozoic thrust
faults (Hutchinson et al., 1986).

BASIN FILL
The Georges Bank sedimentary fill is a prism, thickest over several
small, irregularly shaped, interconnected basins (Figure 2); it thins
gradually seaward of the paleoshelf edge and rapidly toward the platforms
at the hinge zone. The fill is internally divided top to bottom by a

conspicuous unconformity that separates synrift marine and nonmarine


deposits from postrift marine deposits; this unconformity is called the
postrift unconformity (see line 19, Figure 3). The sedimentary prism is also
divided laterally by a buried shelf-edge carbonate platform of Jurassic and
Early Cretaceous age that underlies the present continental slope.
The postrift (or breakup) unconformity (Falvey, 1974; Montadert et al.,
1979) is a conspicuous acoustic reflector that marks a major change in
deposition pattern during basin evolution. This unconformity separates
synrift sedimentary rock from overlying postrift sedimentary rock.
Sediments that accumulated as continental rifting was taking place
(synrift deposits) are 400 m of basal red-to-pink poorly sorted Late Triassic
nonmarine conglomerates and sandstones, and thin marginal marine
limestones, the COST G-1 well (Amato and Bebout, 1980; Arthur, 1982;
Poag, 1982); the redbeds resemble rocks associated with Triassic and
Jurassic grabens onshore (Cornet, 1977; Manspeizer et al., 1978).
The change to a regime of broad basin subsidence was attended by the
formation of a broad carbonate bank, as open marine conditions became
more prevalent. The character of the early postrift deposits changes from a
dominantly carbonate section in the east. Postrift Cretaceous rocks on
Georges Bank are less than half as thick as the postrift Jurassic section (in
COST G-1, about 1300 m versus 2500 m; and in COST G-2, about 1450 m
versus 3900 m; see Figure 2). Subsidence on the margin slowed during the
Cretaceous and Tertiary and became more uniform over the entire margin,
so that little thickening took place. Further, the Cretaceous marked a
change-over to dominantly noncarbonate marine sedimentation.
Continental Rise Sediments
A continental-rise-type feature did not exist during the rift phase of
margin development. The earliest sediments deposited in the new Atlantic
basin were probably continental red beds and evaporites. Accumulation of
these nonmarine and shallow-water sediments continued into the Early
Jurassic late-rift phase of margin development, allowing salt to be deposited
on thinned continental (rift-stage) crust (Folger et al., 1979; Uchupi and
Austin, 1979).

USGS line 19 across the Georges Bank basin

81

Table 1. Inferred reflector ages.

Reflector

Approximate Age

7
9
12
19
23

Top of Cretaceous
Within Cenomanian (mid-Cretaceous)
Top of Jurassic
Hettagian(?) (postrift unconformity)
Pre-Late Triassic

Three conspicuous acoustic reflectors are identified within the Jurassic


deep-sea sedimentary rocks and traced southward toward Deep Sea Drilling
Project (DSDP) sites (Klitgord and Grow, 1980). These reflectors
correspond approximately to the top of the Upper Jurassic (J1), top of the
Middle Jurassic (J2), and within the lower-Middle Jurassic (J3), although
only horizons J1 and J2 are penetrated at DSDP sites.

SUMMARY
The Georges Bank basin shows many similarities in its tectonic setting,
basement structure, and sedimentary fill to other marginal sedimentary
basins along eastern North America (Schlee and Jansa, 1981; Grow and
Sheridan, 1981). Like the adjacent Scotian margin, it is built over a
complexly faulted basement (see Line 19, Figure 3), whose vertical
movement during the early stages of basin formation probably led to synrift
erosion landward of the hinge zone, and to the formation of synrift
depocenters seaward of the hinge zone. This synrift basement structure later
influenced sedimentary-facies and deposition patterns as the Atlantic Ocean
basin formed (Eliuk, 1978; Klitgord et al., 1982; Hutchinson et al., 1986).
In common with other eastern North American marginal basins, Georges
Bank basin formation spans the interval from late Triassic(?) to the present.

REFERENCES CITED

Figure 1. Tectonic elements for the Georges Bank-Gulf of Maine region, including
faults and graben structures on the platforms and along the landward edge of the
Georges Bank basin, location of line 19, salt-diapir province along the seaward edge
of Georges Bank and fracture zones (FZ) and scarps seaward of the bank.

K.D. Klitgord, J.S. Schlee and J.A. Grow

Figure 2. Isopach map of Late Triassic and younger sedimentary rocks in the
Georges Bank basin plus the locations of the two COST wells. Thickness in
kilometers dots show where thickness was measured along the profiles and
indicates the control.

Amato, R.V., and J.W. Bebout, eds., 1980, Geologic operational summary, COST
No. G-1 well, Georges Bank area, North Atlantic OCS: U.S. Geological
Survey Open File Report 80-268, 112 p.
Arthur, M.A., 1982, Lithology and petrography of the COST Nos. G-1 and G-2
wells, in P.A. Scholle, ed., Geological studies of the COST Nos. G-1 and
G-2 wells, United States North Atlantic outer continental shelf: U.S.
Geological Survey Circular 861, p. 11-13.
Cornet, B., 1977, Palynostratigraphy and age of the Newark Supergroup:
Pennsylvania State University, Unpublished PhD Thesis, 506 p.
Eliuk, L.S., 1978, the Abenaki Formation, Nova Scotia shelf, Canadaa
depositional and diagenetic model for a Mesozoic carbonate platform:
Bulletin Canadian Petroleum Geology, v. 26, no. 4, p. 424-514.

USGS line 19 across the Georges Bank basin

82

Figure 3. Seismic line 19, shot northwest to southeast (see Figure 1 for location) across Georges Bank basin. Key reflectors are marked and inferred ages are:
reflector 7, top of the Cretaceous; reflector 9, within the Cenomanian section (mid-Cretaceous); reflector 12, top of the Jurassic; reflector 19, Hettagian(?) (a
postrift unconformity); and reflector 23, post-Early Jurassic(?).

Emery, K.O., and E. Uchupi, 1972, Western North Atlantic Oceantopography,


rocks, structure, water, life, and sediments: AAPG Memoir 17, 532 p.
Emery, K.O., and E. Uchupi, 1984, The geology of the Atlantic Ocean: New York,
Springer-Verlag, 1050 p.
Falvey, D.A., 1974, The development of continental margins in plate tectonic
theory: Australian Petroleum Exploration Association Journal, v. 14, p. 95106.
Folger, D.W., W.P. Dillon, J.A. Grow, K.D. Klitgord, and J.S. Schlee, Evolution of
the Atlantic continental margin of the United States, in M. Talwani et al.,
eds., Deep drilling results in the Atlantic Oceancontinental margins and
paleoenvironment: American Geophysical Union, Maurice Ewing Series 3,
p. 87-108.
Grow, J.A., and R.E. Sheridan, 1981, Deep structure and evolution of the
continental margin off Eastern United States, in Proceedings of the 26th
International Geological Congress, Geology of continental margins

K.D. Klitgord, J.S. Schlee and J.A. Grow

symposium, Paris, 1980: Oceanologica Acta, No. SP, p. 11-19.


Grow, J.A., C.O. Bowin, and D.R. Hutchinson, 1979, The gravity field of the U.S.
Atlantic continental margin: Tectonophysics, v. 59, p. 2752.
Hutchinson, D.R., K.D. Klitgord, and R.S. Detrick, 1986, Rift basins of the Long
Island platform: Geological Society America Bulletin, v. 97, p. 688-702.
Klitgord, K.D., and J.A. Grow, 1980, Jurassic seismic stratigraphy and basement
structure of western North Atlantic magnetic quiet zone: AAPG Bulletin,
v. 64, p. 1658-1680.
Klitgord, K.D., J.S. Schlee, and K. Hinz, 1982, Basement structure, sedimentation,
and tectonic history of the Georges Bank basin, in P.A. Scholle, ed.,
United States North Atlantic outer continental shelf; geological studies of
the COST Nos. G-1 and G-2 wells: U.S. Geological Survey circular 861,
p. 160-186.
Manspeizer, W., J.H. Puffer, and H.L. Cousminer, 1978, Separation of Morocco
and eastern North America-a Triassic-Liassic stratigraphic record:

Geological Society America Bulletin, v. 89, p. 901-920.


Montadert, L., O. de Charpal, D. Roberts, P. Guennox, and J.S. Sibuet, 1979,
Northeast Atlantic passive continental margins; rifting and subsidence
processes, in M. Talwani et al., eds., Deep drilling results in the Atlantic
Oceancontinental margins and paleoenvironment: American
Geophysical Union, Maurice Ewing Series 3, p. 154-186.
Poag, C.W., 1982, Foraminiferal and seismic stratigraphy, paleoenvironments, and
depositional cycles in the Georges Bank basin, in P.A. Scholle, ed.,
Geological studies of the COST Nos. G-1 and G-2 wells, United States
North Atlantic outer continental shelf: U.S. Geological Survey Circular
861, p. 43-92.
Sawyer, D.S., B.A. Swift, J.G. Sclater, and M.N. Toksoz, 1982, Extensional model
for the subsidence of the northern United States Atlantic continental
margin: Geology, v. 10, p. 134-140.
Sawyer, D.S., M.N. Toksoz, J.G. Sclater, and B.A. Swift, 1983, Thermal evolution

of the Baltimore Canyon Trough and Georges Bank basin: AAPG Memoir
34, p. 743-762.
Schlee, J.S., and L.F. Jansa, 1981, The paleoenvironment and development of the
eastern North American continental margin: Oceanologica Acta, 26th
International Congress, Geology of Continental margins, No. SP, p. 71-80.
Sheridan, R.E., J.A. Grow, J.C. Behrendt, and K.C. Bayer, 1979, Seismic
refraction study of the continental edge off the eastern United States:
Tectonophysics, v. 59, p. 1-26.
Swift, B.A., D.S. Sawyer, J.A. Grow, and K.D. Klitgord, 1987, Subsidence, crustal
structure, and thermal evolution of Georges Bank basin: AAPG Bulletin, v.
71, no. 6, p. 702-718.
Uchupi, E., and J.A. Austin, 1979, The geologic history of the passive margin off
New England and the Canadian maritime provinces: Tectonophysics, v. 59,
p. 53-69.

USGS line 19 across the Georges Bank basin

83

CONTOURITES AND VOLCANICS,


GEORGES BANK, NEW ENGLAND

D. BRADFORD MACURDA, JR.


The Energists
Houston, Texas
North America is a classic example of a passive, or trailing, margin. A
strike line from the northeastern United States in the region of Georges
Bank illustrates the deep-water sediments of this area (Figure 1). Basement
(1) occurs at 8.0 sec. In shallower water there is an extensive Jurassic
carbonate bank; the equivalent deep-water carbonate facies is seen from 6.8
to 8.0 sec. (There is a water-bottom multiple just below 7.0 sec that must
not be confused with a sequence boundary.)
Between 6.0 and 6.8 to 6.9 sec, there is a prominent interval with large
mounds. These (at first) resemble giant ripples, climbing to the east. Schlee
et al. (1985) suggest that they are Lower Cretaceous siliciclastic contourite
deposits. Note the large mound in the left center. It shows erosional
t ru n c at ion on its west end just above (2) and downlap at its east end just
above (3). This mound is part of an extensive contourite field, over 300 km
from west to east.
Because the Cretaceous lacked the very cold bottom waters we find
today, warm saline bottom waters probably provided the necessary currents.
A counterpart situation on a smaller scale is seen in the effluent plume of
saline Mediterranean bottom water, which forms contourites off the coast of
Spain and Portugal after crossing the sill at Gibraltar (Faugeres et al., 1984).
One prominent bathymetric feature of the northwestern Atlantic is the
chain of the New England Seamounts. They are presumed to have formed
above a thermal plume in the mantle and to record the movement of the
North American plate. One of these, the Bear Seamount, is shown in Figure
2; it was formed in the Upper Cretaceous (Duncan, 1984). The top of the
Lower Cretaceous contourites is just above 6 sec; a wedge of volcaniclastic
debris is seen mantling above it (4). (The reflection surface at (5) is a
multiple.) Tertiary siliciclastics onlap the seamounts. There is a prominent
erosional unconformity formed during the Oligocene, with erosional
truncation at the edge of the submarine canyon, seen above (6).
Other basement highs occur in proximity to the seamounts. The high in
Figure 3 occurs a short distance to the southwest of the Bear Seamount. It

D.B. Macurda, Jr.

cuts through the Jurassic and Lower Cretaceous sediments, and may be the
eroded remnants of a volcano.
The record of the Tertiary is complex, with several intervals of
deposition and erosion. The unit labeled (7) in Figure 1 is Paleocene-Eocene
in age; the lens-shaped unit with the hummocky reflectors just above and to
the right of (7) is a large mass-debris flow that occurred within this interval.
One of the most interesting aspects of the uppermost part of the section is
the occurrence of numerous closely-spaced faults (see reflector marked 8).
These probably record the dewatering of the sediments during compaction.
These features are extensively developed in deep-water basin-fill facies in
other basins, such as the Navarin basin in the Bering Sea, Gulf of Mexico,
and North Sea.

REFERENCES CITED
Duncan, R.A., 1984, Age progressive volcanism in the New England Seamounts
and the opening of the North Atlantic: Journal Geophysical Research, v.
89, p. 9980-9990.
Faugeres, J.C., E. Gonthier, and D.A.V. Stow, 1984, Contourite drift molded deep
Mediterranean outflow: Geology v. 12, no. 5, p. 296-300.
Schlee, J.S., C.W. Poag, and K. Hinz, 1985, Seismic stratigraphic of the
continental slope and rise seaward of Georges Bank, in C.W. Poag, ed.,
Geologic evolution of the United States Atlantic margin: New York, Van
Nostrand Reinhold Co., p. 265-292.

Contourites and volcanics, Georges Bank

84

SHELF-TO-BASIN CORRELATIONS OFF EASTERN CANADA: DEVELOPING A


SEISMIC STRATIGRAPHIC FRAMEWORK IN THE NORTHERN NEWFOUNDLAND BASIN

KAREN J. MEADOR
JAMES A. AUSTIN, JR.
and
DONALD F. DEAN
University of Texas, Institute for Geophysics
Austin, Texas
INTRODUCTION
We have been able to interpret structural features, locate the oceanic/
continental-crust boundary, and establish a stratigraphic framework for the
northern Newfoundland basin off eastern Canada (Figure 1) by using three
seismic stratigraphic techniques:
1. Calibration of lithology and velocity logs from a shelf well, the
AMOCO-Imperial Oil-Skelly Skua E-41, to the established seismic
stratigraphy of the Grand Banks,
2. Correlation of U, a regionally prominent acoustic unconformity,
from the shelf to the adjacent deep basin using newly acquired
multichannel seismic (MCS) and magnetic field data, and
3. Comparison of the interpreted synrift/early drift stratigraphy of the
northern Newfoundland basin with the coeval section on the conjugate
passive margin off the Iberian peninsula.
The MCS lines and associated magnetic field data presented here are the
results of a survey (termed NB in this chapter) conducted in the
Newfoundland basin during August and September, 1984, by the University
of Texas Institute for Geophysics and the Woods Hole Oceanographic
Institution aboard the Lamont-Doherty Geological Observatory research
vessel Conrad (Figure 1).

NB-2/SKUA E-41 TIE


The initial step in developing a regional seismic stratigraphy for the
northern Newfoundland basin was to tie lithologies from the Skua E-41

K.J. Meador, J.A. Austin, Jr. and D.F. Dean

shelf well to MCS line NB-2, which purposely crossed the well bore
(Figure 1). The lithology log (Figure 2a) and velocity log (Figure 2b,
converted from the sonic log) from the well exhibited the following
significant features: both lithologic and velocity cycles in the uppermost
Jurassic (J)-Cretaceous (K) section (Figures 2a, 2b); a coarse sandstone at
the top of the Lower Cretaceous (depth approximately 1.31 km or 4292 ft);
and a prominent limestone interval at the top of the Upper Cretaceous
(depth approximately 1.1 km or 3602 ft). First, an acoustic impedance log
was generated using the velocity and density logs, which were handdigitized and resampled at 0.62-m (2-ft) intervals. Then, a normal-incidence
reflectivity log (Figure 2c) in units of both travel-time and sub-sea-level
depth was constructed using the Digicon DISCO Wavelet Processing
package (see Acknowledgments). Next, normal-incidence velocities from a
downhole check shot survey of the Skua E-41, obtained from AMOCO,
were converted to two-way travel-times for the raypaths characteristic of the
seismic acquisition system on Conrad (see captions for Figures 3 and 4). Finally,
these times were used as guides to picking NB-2 stacking velocities (Figure 3).
The resultant stack of the part of NB-2 that straddles the Skua E-41 well
site could be accurately tied to the lithologic section sampled by the well
(Figure 4). Generally, reflections from the uppermost J-K section correlate
with limestone/sandstone/shale cycles (Figure 2a) and their corresponding
cyclic velocity variations (Figure 2b). For example, the high-amplitude,
continuous reflector at 1.15 sec on NB-2 at the well-bore (Figures 3 and 4)
correlates with the top of the Upper Cretaceous limestone sequence
previously mentioned (Figure 2a), which in turn corresponds both to the
largest velocity increase encountered in the Skua E-41 well (Figure 2b) and
to a pronounced peak in reflectivity (Figure 2c). We can tie this highamplitude, continuous-reflecting surface to the ubiquitous U
unconformity (AMOCO and IMPERIAL, 1973; Grant, 1977) on the basis
of their acoustic similarities. However, this acoustic surface is too young to
be the late Early Cretaceous unconformity that U is interpreted to
represent on the Grand Banks (AMOCO and IMPERIAL, 1973).
Geologically, U must correlate instead with the coarse sandstone that
constitutes the top of the Lower Cretaceous section in the Skua E-41
(Figures 2a and 3). This lithologic contrast occurs at approximately 1.29 sec
on NB-2 (Figure 4), and while it correlates with another velocity

discontinuity and reflectivity peak in the well (Figures 2b and 2c), it does
not generate the large impedance contrast on NB-2 that acoustic U does
immediately above it. Nonetheless, inspection of nearby NB (Figure 1) and
industry MCS lines (Hubbard et al., 1985; Meador, in preparation) have
convinced us that while U can indeed be identified on the northeastern
part of the Grand Banks as package of high-amplitude, parallel reflections
that generally truncate underlying structures (AMOCO and IMPERIAL,
1973), the amplitude of individual reflectors within that package can vary
substantially over distances as short as a few kilometers. In fact, the
package of parallel reflections that comprises U on the shelf appears to
converge toward the shelf-break (Meador, in preparation), and only the
sequence boundary representing geologic U (the presumed late Early
Cretaceous hiatus separating the coarse sandstone and overlying shale and
limestone units) continues into the adjacent, deep basin. Previous attempts
to trace U into the adjacent basins have been inconclusive, either because
of the disruptive nature of steep slope topography or because of
inadequacies in the quantity or quality of the geophysical data used to make
the correlations (e.g., Grant, 1977, 1979; Parson et al., 1985).
Apparently, the reflection generated by the top of the Upper Cretaceous
limestone, acoustic U at the Skua E-41, persists only in shallow water,
coincident with the regional extent of a shelf carbonate facies. This
underscores the importance of tying borehole data to all available regional
seismic lines, because locally prominent facies contrasts can produce large
acoustic impedances, effectively masking more widespread and
geologically significant unconformities.

SHELF-TO-BASIN CORRELATION OF U AND THE


CRUSTAL EVOLUTION OF THE NORTHERN
NEWFOUNDLAND BASIN
Figure 5 shows a map compilation of magnetic anomalies in the
Newfoundland basin, projected along NB survey tracks (see also Figure 1).
Note that on all lines north of the Southeast Newfoundland Ridge, the
residual magnetic field exhibits a dual character. Over the outer part of the
Grand Banks and the inner half of the Newfoundland basin, the field is
relatively smooth. However, along a trend approximated by the 4000-m

bathymetric contour, the amplitude of the residual field abruptly increases.


Keen et al. (1977) and Sullivan (1983) have interpreted this multi-peaked
positive anomaly in the Newfoundland basin south of the Newfoundland
Seamounts as the northern expression of the J-anomaly, a late Early
Cretaceous (M0-M1) sea-floor-spreading lineation originally identified and
mapped south of the Southeast Newfoundland Ridge (Pitman and Talwani,
1972; Rabinowitz et al., 1978, 1979) (Figure 1). Both Sullivan (1983) and
Masson and Miles (1984) have interpreted the J-anomaly as the oldest seafloor-spreading anomaly in the Newfoundland basin, and have therefore
used it and its mapped equivalent off Portugal (Group Galice, 1979) to
reconstruct the pre-drift configuration of this part of the North Atlantic
(LePichon et al., 1977). While the magnetic character of the Newfoundland
basin as revealed by the NB anomaly profiles would appear to support such
interpretations, the NB survey results indicate for the first time that the Janomaly also can be traced (perhaps with slight eastward offset) north of the
Newfoundland Seamounts (Figure 5) (Meador, in preparation).
As previously stated, a late Early Cretaceous unconformity/geologic
U can be traced from the Skua E-41 well site across the shelf and slope to
the northern part of the deep Newfoundland basin, by using both the NB
survey results (NB lines 1 and 3, Figure 1) and available industry MCS lines
(Meador, in preparation). On the slope, U steepens and truncates
underlying reflectors at low angles. Overlying reflectors onlap. In the inner
part of the deep basin, U forms the top of a reverberant sequence of
reflectors that onlap and infill acoustic basement. U is continuous and
generally convergent with the tops of basement blocks, which appear flattopped and steep-sided (faulted?) and exhibit little or no magnetic signature.
Farther seaward, U onlaps basement flanks before pinching out in the
vicinity of the J-anomaly (Figures 6 to 8).
In this part of the Newfoundland basin, the J-anomaly pinpoints a major
change in the morphology of basement (Figures 6 to 8). Beneath the Janomaly and continuing on to the east, basement is both rougher and
shallower than basement to the west (Figures 6 to 8). We interpret this
change in crustal morphology, which mirrors the major change in the
residual magnetic field, to mark the boundary between rifted continental
crust to the west and oceanic crust to the east in the deep, northern
Newfoundland basin. The pinch-out of the U unconformity at this crustal
boundary corroborates that U is late Early Cretaceous (Barremian-Aptian,

Shelf-to-basin correlations off eastern Canada

88

Figure 1. Bathymetric map of the Newfoundland basin showing location of the NB


seismic grid. Portions of seismic profiles NB-2 (Figure 4) and NB-4 (Figures 6 and 7)
are highlighted with a bold line. Note the location of the Skua E-41 industry well on
NB-2. The migrated portion of NB-4 (Figure 8) is shown as a dashed line. Inset map
shows the region of the southeastern Grand Banks shelf/eastern Canadian passive
continental margin.

approximately the same age as the J-anomaly), and further implies that U
marks the rift-drift transition on this part of the eastern Canadian passive
continental margin. If this preliminary interpretation of basement in the
northern Newfoundland basin is correct, then rifted continental material
extends seaward almost 400 km from the edge of the Grand Banks in this
region.

SEISMIC STRATIGRAPHY: DEEP NORTHERN


NEWFOUNDLAND BASIN
Line NB-4 illustrates that the northern Newfoundland basin
stratigraphic succession is characterized by five major seismic sequences,
S1-S5 (Figures 6 to 8). S1 lies below the U unconformity and infills
basement lows west of the J-anomaly. S1 is highly variable in thickness and
is absent seaward (or east) of the J-anomaly. Because we have identified
U as the rift-drift transition on this passive margin, S1 must represent a
synrift deposit composed of sediments eroded from adjacent continental
blocks and/or volcanics associated with rifting. The presence of
volcanics/volcaniclastics might explain the high amplitudes of reflectors
within this sequence, but the lack of large magnetic anomalies over this part
of the basin argues against a significant volcanic component in the synrift
section.
S2 is the earliest drift sequence immediately overlying the U
unconformity (Figures 6 to 8). This sequence is composed of moderately
continuous, low-to-medium-amplitude reflections. S2 smooths most of the
remaining topography created by underlying (rifted continental) basement
west of the J-anomaly, and constitutes the oldest sediments immediately
overlying (oceanic) basement east of the J-anomaly. S2 is equivalent to
sequence 4 of Parson et al. (1985), which they tentatively correlate with
clastics associated with latest Early Cretaceous uplift and erosion of the
then still-emergent Grand Banks.
Sequence S3 can be divided into two subsequences, S3a and S3b
(Figure 7). S3a is composed of high-amplitude, continuous, parallel
reflections that become inclined east of the J-anomaly. This seismic facies is
generally characteristic of hemipelagic deposition in a low-energy
environment (Vail et al., 1977). S3a maintains a constant thickness
throughout the deep basin, and pinches out on the mid-to-lower part of the
slope (Meador, in preparation).
S3b is composed of moderately continuous to discontinuous, medium-

K.J. Meador, J.A. Austin, Jr. and D.F. Dean

amplitude reflectors that downlap the top of S3a. A hummocky facies


intercalated with flat-lying, continuous reflections becomes more prevalent
toward the slope, suggesting the presence of gravity-controlled clastic
processes there, i.e., debris flows, proximal turbidites (Vail et al., 1977;
Brown and Fisher, 1980). S3b thickens toward the slope, and pinches out
slightly updip from the pinchout of S3a. S3 can be correlated with the lower
part of sequence 3 of Parson et al. (1985), which they ascribe to continued
early Late Cretaceous clastic black shale input to the Newfoundland
basin.
S4 is a thick sequence of variable-amplitude, prograding reflections; it
systematically downlaps the top of S3. Seismic facies suggest that
deposition of S4 was predominantly controlled by deep-water processes. As
with S3, S4 can be divided into two subsequences, S4a and S4b (Figure 7).
S4a is roughly equivalent to the upper part of sequence 3 of Parson et al.
(1985), while S4b is generally equivalent to their sequence 2, interpreted as
a Late Cretaceous-early Tertiary sequence of interbedded chalks and
mudstones. This part of S4 near the slope may also be stratigraphically
equivalent to the Flemish Cap wedge (Sullivan, 1983).
The top of S4 is abruptly truncated by the lower boundary of S5, the
uppermost sequence identified in the northern Newfoundland basin (Figures
6 to 8), which may at least in part be related to an intensification of deep,
western North Atlantic circulation patterns associated with northern
hemisphere cooling in the late Eocene to early Oligocene (Miller and
Tucholke, 1983). S5 is composed of generally high-amplitude, variable to
continuous, wavy reflectors that exhibit mounded to chaotic structures and
basal scour surfaces. Sullivan (1983) has interpreted this section as
turbidites, but the seismic facies suggest that S5 deposition has been
controlled by contour-following bottom currents, which appear to have been
at work in this region since the early Miocene (Miller and Tucholke, 1983;
Parson et al., 1985).

PRE-DRIFT RECONSTRUCTION VS.


STRATIGRAPHIC EVOLUTION
A pre-drift plate reconstruction of the northern North Atlantic suggests
that this part of the Atlantic basin was narrow at J-anomaly time (LePichon
et al., 1977; Meador, in preparation). Consequently, the rift/early drift
stratigraphic successions sampled by the Skua E-41 well and DSDP Site
398 drilled off Portugal should be generally comparable. Based upon a
comparison of MCS profiles from both basins (Meador, in preparation), we
can correlate the U unconformity to the coeval 3/4 sequence boundary, an
Albian/Aptian surface, off the Iberian peninsula (Group Galice, 1979). Both
S1 and sequence 4 (Group Galice, 1979) are moderately to strongly layered,
infill basement lows, and exhibit highly variable thicknesses. Both S2 and
sequence 3 (Group Galice, 1979) are composed of lower-amplitude, less

Shelf-to-basin correlations off eastern Canada

89

Figure 5. Magnetic anomalies plotted along dip lines of the NB survey. Shaded area is
positive. Approximate trend of the J magnetic anomaly in the northern
Newfoundland basin is highlighted. The J-anomaly is also plotted on Figures 6-8
above seismic line NB-4.

continuous reflectors. Such comparisons provide both additional


corroboration of our shelf-to-basin correlation of U and underscore its
stratigraphic significance as the rift-drift transition on this part of the
Canadian passive continental margin.

ACKNOWLEDGMENTS
Lithologies shown in Figures 2 to 4 are based on AMOCO-Imperial OilSkelly well logs, which were sent to us by Rileys Data Service, Calgary,
Alberta, Canada. Downhill velocity information for the Skua E-41 well was
taken from a check shot survey run for AMOCO Canada Petroleum
Company, Ltd. Results from this survey were made available to us through
the help of the Canada Oil and Gas Lands Administration (COGLA),
Ottawa, Ontario, Canada. Ages for Skua E-41 lithologies are based on
micropalentological analyses performed by AMOCO Production Research
Center, Tulsa, Oklahoma.
All seismic and magnetic field data collected during the NB-survey have
been processed at the University of Texas Institute for Geophysics, using
DISCO processing software developed and marketed by Digicon, Inc.,
Houston, Texas. We are indebted to Malcolm I. Ross for helping to generate
the magnetic anomaly profiles.
We especially thank Dr. Brian E. Tulcholke of Woods Hole
Oceanographic Institution, co-principal investigator of the NB-project and
co-chief scientist during the Conrad survey, for his continuing input and
encouragement.
Funding for this work was provided jointly by the Ocean Sciences
Division and the Ocean Drilling Program of the National Science
Foundation through Grant Not. OCE-8308623.
This chapter constitutes University of Texas Institute for Geophysics
contribution no. 659.

Newfoundland Ridge: Tectonophysics, v. 59, p. 71-81.


Group Galice, 1979, The continental margin off Galicia and Portugal: Acoustical
stratigraphy, dredge stratigraphy, and structural evolution: Initial Reports
of the Deep Sea Drilling Project, v. 47B, p. 633-662.
Hubbard, R.J., J. Pape, and D.G. Roberts, 1985, Depositional sequence mapping to
illustrate the evolution of a passive continental margin: AAPG Memoir 39,
p. 92-115.
Keen, C.E., B.R. Hall, and K.D. Sullivan, 1977, Mesozoic evolution of the
Newfoundland basin: Earth and Planetary Science Letters, v. 37, p. 307320.
LePichon, Z., J.C. Sibuet, and J. Francheteau, 1977, The fit of the continents
around the North Atlantic Ocean: Tectonophysics, v. 28, p. 169-209.
Masson, D.G., and P.R. Miles, 1984, Mesozoic seafloor spreading between Iberia,
Europe and North America: Marine Geology, v. 56, p. 279-287.
Meador, K.M., in preparation, Seismic stratigraphic framework of the northern
Newfoundland basin: M.A. thesis, The University of Texas at Austin.
Miller, K.G., and B.E. Tucholke, 1983, Development of Cenozoic abyssal
circulation south of Greenland-Scotland Ridge, in B. Bott, S. Sapor, M.
Talwani, and J. Thiede, eds., Structure and development of the GreenlandScotland Ridge: Plenum Press, New York, p. 549-589.
Parson, L.M., D.G. Masson, C.D. Pelton, and A.C. Grant, 1985, Seismic
stratigraphy and structure of the east Canadian continental margin between
41 and 52N: Canadian Journal of Earth Sciences, v. 22, p. 686-703.
Pitman, W.C., III, and M. Talwani, 1972, Sea-floor spreading in the North Atlantic:
GSA Bulletin, v. 83, p. 619-646.
Rabinowitz, P.D., S. Cande, and D.E. Hayes, 1978, Grand Banks and J-anomaly
Ridge: Science, v. 202, p. 71-73.
Rabinowitz, P.D., S. Cande, and D.E. Hayes, 1979, the J-anomaly in the central
North Atlantic Ocean: Initial Reports of the Deep Sea Drilling Project, v.
43, p. 879-885.
Sullivan, K.D., 1983, The Newfoundland basin: ocean-continent boundary and
Mesozoic seafloor spreading history: Earth and Planetary Science Letters,
v. 62, p. 321-339.
Vail, P.R., R.M. Mitchum, Jr., R.G. Todd, J.M. Widmier, S. Thompson, III, J.B.
Sangree, J.N. Bubb, and W.G. Hatlelid, 1977, Seismic stratigraphy and
global changes of sea level: AAPG Memoir 26, p. 49-212.

REFERENCES CITED
AMOCO Canada Petroleum Company, Ltd., and IMPERIAL Oil Ltd., 1973,
Regional geology of the Grand Banks: Bulletin of Canadian Petroleum
Geology, v. 21, p. 479-503.
Brown, L.F., Jr., and W.L. Fisher, 1980, Seismic stratigraphic analysis
interpretation and petroleum exploration: AAPG Continuing Education
Course Notes 16, p. 1-181.
Grant, A.C., 1977, Multichannel seismic reflection profiles of the continental crust
beneath the Newfoundland Ridge: Nature, v. 270, p. 22-25.
Grant, A.C., 1979, Geophysical observations bearing upon the origin of the

K.J. Meador, J.A. Austin, Jr. and D.F. Dean

Shelf-to-basin correlations off eastern Canada

92

AGGRADING AND PROGRADING INFILL OF BURIED CENOZOIC SEAWAYS,


NORTHWESTERN GREAT BAHAMA BANK

GREGOR P. EBERLI
and
ROBERT N. GINSBURG
University of Miami
Rosenstiel School of Marine and Atmospheric Science
Comparative Sedimentology Laboratory, Fisher Island Station
Miami Beach, Florida
INTRODUCTION
The first good multichannel seismic profiles across the top of
Northwestern Great Bahama Bank reveal that this part of the platform has
not grown up continuously with the configuration it now has, but is instead
the result of the coalescence of three smaller platforms and their lateral
expansion (Eberli and Ginsburg, 1987). The coalescence resulted from
infilling of the seaways by aggrading and prograding systems. The grid of
high-quality profiles makes it possible to suggest a mechanism to explain
the origin of the seaways and to delineate the anatomy of their fills.
Recognition of large-scale progradation capable of closing seaways as large
as the present-day Tongue of the Ocean in order to connect separate
platforms is an alternative model of platform evolution. The earlier view
was that carbonate slopes steepen with time and consequently a growing
platform shrinks in size (Schlager and Ginsburg, 1981).
The seismic data available for the present study consist of the top parts
(1.1 to 1.7 sec two-way travel time) of mostly unmigrated multichannel
profiles provided by Texaco Inc. and Western Geophysical. Figure 1 gives
the location of the profiles. The grid of profiles connects with the
exploratory well Great Isaac-1, and its velocity profile is the base of the
depth conversion and the reported ages for stratigraphic interpretations
(Tator and Hatfield, 1975; Schlager et al., in press).

THE BURIED SEAWAYS


From the grid of seismic profiles, two north-south-trending depressions
are recognized in the subsurface of northwestern Great Bahama Bank. They

G. Eberli and R. Ginsburg

differ in size, age, and internal structure, but were probably both initiated by
tectonic deformation (Eberli and Ginsburg, in press).
An older, eastern depression, termed the Straits of Andros, separated the
Andros Bank to the east from the Bimini Bank to the west (Figures 2 and
3). The straits, approximately 25 km wide, can be followed for about 70 km
without a change in dimensions. Three features of the seismic profiles
suggest that the origin of the straits was fault-controlled. First, there is an
abrupt termination of the lowermost reflections in the straits against the
west side of the Andros Bank. Second, there are refractions within the
Andros Bank that are interpreted as the result of faults; and third, there is
the inferred offset of a horizon of high-amplitude reflections between the
Banks and the straits. The correlation horizon of Figure 3 is believed to be
the same as the horizon marked c at the bottom of the straits. This
interpretation means a displacement of 1500 m that originated the straits.
The age of this prominent reflector can be inferred from the dated section in
the Great Isaac exploration test. In this well, a similar high-amplitude
horizon marks the top of a mid-Cretaceous carbonate platform. If this
interpretation is correct, the straits of Andros were initiated in the mid-Cretaceous.
In middle to late Tertiary time, the second, younger depression, Bimini
Embayment, formed within the Bimini Bank by the combination of tectonic
deformation and differential sedimentation. Deformation is seen in a largescale fold with an amplitude of 300 m and small-scale faults under the
depression. Differential sedimentation is recognized on the east side of the
embayment by a wedge of downlapping reflections (labeled with an arrow,
Figure 5b). The embayment, approximately 470 m deep and 10 km wide,
narrowed and shallowed to the north and was probably not connected most
of the time with the Northwest Providence Channel, but was open to the
already existing and then broader Straits of Florida (Figure 2).

THE FILLING OF THE SEAWAYS


The filling of the seaways was the combined result of aggradation and
significant progradation of the platform margins. These processes filled the
Straits of Andros and the Bimini Embayment completely, and extended the
western margin of Bimini Bank 25 km out into the Straits of Florida to its
present position (Figure 3). Within both straits the filling consists of
aggradation followed by progradation.

Figure 1. Location map of the seismic lines; insert = location of the


study area.

Aggradational Phase
The anatomy of the aggradational phase is clearly revealed in the Straits
of Andros (Figure 4). Initially, the straits were asymmetric in cross section.
This initial asymmetry was leveled by a wedge-shaped fill with a horizontal
upper boundary and internal, continuous, moderately high-amplitude
reflections. Over the nearly horizontal floor a second wedge-shaped, mostly
incoherent prism developed on the east side of the straits. West of this
wedge, continuous reflections run into the axis of the straits and form a

Figure 2. Paleogeographic map of northwestern Great Bahama Bank at midTertiary (?) time, showing the two north-south-trending depressions of the
Straits of Andros and Bimini Embayment. Soundings in msec; stipple = 250300 msec. (From Eberli and Ginsburg, 1987).

relatively thick unit, suggesting that the sedimentation rate in the axis of the
straits was slightly higher than on the slope, and as a consequence the basin
shallowed and the slope angle decreased. A slope of 2 to 4 developed,
which is characterized by high-amplitude reflections and abundant
refractions.
We do not have any direct information on the composition of these
units, but the results of ODP Leg 101 and the Great Isaac well (Austin et al.,
1986 Schlager et al., in press) can be used for a tentative interpretation. The

Buried Cenozoic seaways, northwestern Great Bahama Bank

97

basal unit, filling the initial asymmetry may consist of intercalations of


various proportions of mass gravity flow deposits (debris sheets, carbonate
turbidite and fine-grained (periplatform) ooze and chalk. The second wedge
that is seismically mostly incoherent may consist mainly of coarse, neritic
debris from the adjacent Andros Bank. The continuous reflections west of
the wedge suggest that mass gravity flows were then preferentially
deposited in the axis of the straits that had shifted westward. The low-angle
slope displays characteristics of an accretionary slope (Schlager and
Ginsburg, 1981). Its reflections are interpreted as intercalations of carbonate
turbidites with periplatform ooze, suggesting that on the less inclined slope,
mass gravity flows were able to deposit. The refractions may be caused by
slumped units. Seismic and drilling investigations along the modern openocean slope of Little Bahama Bank revealed that slumping and creeping are
common features along the toe-of-slope of the this similar, accretionary
slope (Mullins et al., 1984; Austin et al., 1987).
Prograding PhaseStraits of Andros

Prograding Phase-Bimini Embayment


In the Bimini Bank, the mid-Cretaceous(?) correlation horizon is
overlain by four megasequences (Figure 5). The top of megasequence B
represents a major unconformity on which the clinoforms of the Bimini
Embayment downlap. The reflection marking the unconformity can be
followed into the Straits of Andros, where it overlies the prograding system.
It is assumed that the formation of the Bimini Embayment is coeval with the
unconformity and is therefore a much younger feature than the Straits of
Andros and the Straits of Florida.
As indicated by the geometry of the sequences filling the embayment,
sediment transport in the Bimini Embayment was, as in the Straits of
Andros, predominantly from east to west. Slope steepness and the ratio of
progradation versus aggradation in this sigmoidal prograding system both
indicate a low-energy environment (Mitchum et al., 1977), which agrees
with the more protected position of the embayment within the Bimini Bank.
The sediments filling the embayment may consist of various proportions of
platform-derived and pelagic components.

CONCLUSIONS
A dramatic change in the filling of the two straits is seen at a depth of 1
sec (two-way travel time), when over the slope an east-west prograding
system develops (Figure 4). In the Straits of Florida, the system develops
from wide, indistinct sigmoids to steep sigmoidal sequences. The modern
platform edge is the latest of these prograding sequences (Figure 3). In the
Straits of Andros the system also starts with three wide, sigmoidal
sequences and develops into a system of complex sigmoid-oblique
sequences (Figure 4a). Each of these sequences is characterized by a highamplitude reflection on the sigmoid front and on its flat upper part.
Reflections are weak or absent over the edge of the sigmoid. These
reflection-free spots are interpreted as reef build-ups, and the steep
reflections in front could represent the successive fore-reef slopes. Flatlying, lagoonal sediments may be responsible for the well-developed highamplitude reflections on the back of the sigmoids. Close to the sigmoid
edge the slopes reach 23, but basinward they flatten rapidly. The elevation
difference from the reef crest to the depression floor is between 400 and
500 m. The total aggradation of the entire prograding system is 300 m; in
the same time it prograded approximately 10 km. This
aggradation/progradation ratio indicates a very rapid progradation; the
sigmoids have a geometry characteristic of a high-energy system (Mitchum
et al., 1977). Sedimentation was not confined to the prograding system, but
continued in the straits, so that the clinoforms onlap and climb over the
aggrading unit, a feature Bosellini (1984) describes for the prograding
Carnian Sella Platform.

G. Eberli and R. Ginsburg

The profiles show that two phases of segmentation have occurred in the
Great Bahama Bank since the mid-Cretaceous. Subsequent coalescence by
aggrading and prograding sequences documents the dynamic quality of the
carbonate environment and the influence of lateral extension in the platform
evolution.
The presence of such distinct reflections in these young, entirely
carbonate deposits is surprising. The anatomies of the aggrading and
especially the prograding systems are remarkably similar to siliciclastic
sequences.
The change from aggradation to progradation is likely the result of
changes in the relative position of sea levelhighstands for the aggrading
phase and low stands for the progradation.

ACKNOWLEDGMENTS
We thank Texaco, Inc. and Western Geophysical for providing the
profiles. We appreciate the encouragement and helpful discussions with
Robert M. Galbraith and colleagues of Texacos Coral Gables Office. We
acknowledge support from Industrial Associates of the Comparative
Sedimentology Laboratory.

REFERENCES CITED
Austin, J.A., Jr., et al., 1986, Proceedings of the Ocean Drilling Program,
Bahamas, Covering Leg 101 Initial Report: Washington, D.C., U.S.
Government Printing office, 247 p.
Bosellini, A., 1984, Progradation geometries of carbonate platforms: examples
from the Triassic of the Dolomites, northern Italy:Sedimentology, v. 31, p. 1-24.
Bosellini, A., 1988, Outcrop models for seismic stratigraphy: examples from the
Triassic of the Dolomites: this volume.
Eberli, G.P., and R.N. Ginsburg, 1987, Sedimentation and coalescence of Cenozoic
carbonate platforms, northwestern Great Bahama Bank: Geology, in press.
Mitchum,R.M.,Jr., P.R. Vail, and J.B. Sangree, 1977, Seismic stratigraphy and global changes
in sea level,part 6: Stratigraphic interpretation of seismic reflection patterns in
depositional sequences, in C.E. Payton,ed., Seismic stratigraphyapplications to
hydrocarbon exploration:AAPG Memoir 26, p. 117-133.
Mullins, H.T., et al., 1984, Anatomy of a modern open-ocean carbonate slope:
Northern Little Bahama Bank: Sedimentology, v. 31, p. 141-168.
Schlager, W., and R.N. Ginsburg, 1981, Bahama carbonate platformsthe deep
and the past: Marine Geology, v. 44, p. 1-24.
Schlager, W., in press, Great Isaac well, in Austin et al., Proc. Rept., ODP Leg 101,
Part B, Washington D.C., U.S. Government Printing Office.
Tator, B.A., and L.E. Hatfield, 1975, Bahamas present complex geology: Oil and
Gas Journal, v. 73, no. 43, p. 172-176 and no. 44, p. 120-122.
Appendix 1: Data Information WESTERN line (Figures 3 and 4).
Energy source
Airguns
Number of guns
10
Total gun volume
720 P.S.I.
Firing interval
164 feet
Shot point interval
164 feet
Distance of source to antenna
122 feet
Type cable
drag yo-yo
Average cable depth
16 feet
Processing sampling interval
2 msec F-K filter
Deconvolved before stack
Time variant filter
No migration
Appendix 2. Data information line 1-N-C (Figure 5).
Energy source
Airguns
Number of guns
12
Gun volume
1120 cu3
Gun pressure
4700 P.S.I.
Firing interval
164 feet
Shot point interval
164 feet
Distance of source to antenna
122 feet
Type cable
drag yo-yo
Average cable
depth 14 feet
Processing sampling
F-K filter
Deconvolved before stack
Time variant filter
Migration by the finite difference method

interval 2 msec

Buried Cenozoic seaways, northwestern Great Bahama Bank

98

DEEP CLASTIC CARBONATE DEPOSITS OF THE BAHAMASCOMPARISON WITH


MESOZOIC OUTCROPS OF THE VERCORS AND VOCONTIAN TROUGH

C. Ravenne and R. Vially


IFP
Rueil-Malmaison, France
P. LeQuellec
Compagnie Francaise des Petroles (CFP)
Paris la Defense, France
and
P.Valery
Societe National Elf Aquitaine (SNEA)
Paris la Defense, France
INTRODUCTION
We present some results of a high-resolution seismic survey conducted
at the foot of the Bahama Scarp, in the vicinity of the outlet of the Great
Bahama Canyon (Figure 1) (Ravenne et al., 1985; Ravenne et al., in
preparation). Also, where possible, we compare the seismic data with the
ground data obtained in the Vocontian trough (Figure 2) (Ravenne and
Vially, in preparation).
We are only concerned here with sequences II and III, the Miocene to
Quaternary, whose construction is the result of two distinct gravity
processes. The first, responsible for the construction of the major part of
sequence II, corresponds to mass slidings resulting from the stripping of the
shelf and its slope. The second, responsible for the construction of the major
part of sequence III, involves turbidites and contourites. The material is
drained by a canyon and then by distributary channels.
Mass slides still occur, but the slip material is remobilized by currents
running along the slope. The currents are active in building levees supplied
by deposits from the canyon and from the slipped masses.

C. Ravenne, P. LeQuellec, P. Valery and R. Vially

Our interpretation is highly simplified, in order to highlight sequences II


and III. Only the analysis of profile BAC E12 (Figures 3A-3H) is detailed
and shows how the correlations are established. The many faults affecting
unit II on the slope are not shown (to avoid cluttering the profiles). All
profiles perpendicular to the slope were cut at the point where organized
reflections disappeared.

DESCRIPTION AND EVOLUTION OF SEQUENCE III:


CHANGES IN SEISMIC FACIES UNITS
Profile BAC E14 (2) (Figures 4A-4D) shows the enormous development
of sequence II at the base of the slope. It reveals many internal
unconformities separating different seismic facies units. Only a few of these
unconformities are depicted. On this section, subsequence II b appears to lie
in relative conformity with the underlying sequence. Towards shotpoint 830,
this section intersects profile BAC E8 (Figure 5A; shotpoint 320), where
subsequence II b lies in complete unconformity on the lower terms.
Similarly, subsequence II c completely erodes subsequence II b. In detail,
many more unconformities are distinguished in subsequence II b and II c.
The form of subsequences II b and II c as well as the large number of
internal unconformities clearly show that these subsequences result from a
succession of slipped packages originating in the slope and the shelf. A
close analysis of the seismic facies units of sequence II on profile BAC E
14-2 (Figure 4) helps to identify different shelf and slope elements,
fractured and faulted but still well organized.
This succession of slides is present all along the slope and is reflected in
the morphology by lobate forms on the slope. The maximum deformation
occurs between profiles BAC E12 (Figure 3) and BAC E15 (Figure 6).
These slides are responsible for the eastward inflection of the outer channel
of Great Bahama Canyon.
Subsequence II a (Figure 3) generally displays very poorly organized
reflections and seems to have served as a slip bed for the upper
subsequences. The slipped unit (subsequences II b and II c) as seen on
profile BAC E8 (Figures 5A and 5B) is over 10 km wide and about 1 km
thick. Profile BAC E14-2 (Figure 4) shows that the bedding is still visible
over more than a 15-km distance.

These profiles reveal sedimentary bodies resulting exclusively from


mass slide processes involving large volumes. On the outcrops, only the
carbonate megaturbidites (Johns et al., 1981; Labaume, 1983; Labaume et
al., 1983) involve significant volumes. These megaturbidites consist of beds
containing redeposited carbonate shelf material. Commonly, they exhibit
dimensions compatible with the seismic resolution power. Thus, the
thickness of these bodies may exceed 200 m and their width may reach 10
to 20 km. Moreover, these bodies generally are concordant at the base and
at the top with the surrounding terrigenous turbidites.
The slip of these carbonate bodies seems to have taken place on a fluid
bed, occurring where a gigantic water escape forms and preserves the
substratum from erosion. This fluid bed was identified during Ravenne and
Beghins modeling of turbid flows in a channel (1983). The role of turbid
flows in the slides and the preservation of the substratum from erosion is
discussed, in relation to the deposition of bodies of the granular bar type in
deep environments.
It is highly probable that other sedimentary bodies of the same origin
and of comparable size to that shown on the seismic profiles exist in
outcrops, but these are not commonly described as such because of their
large size (only seismic surveys over a large area can demonstrate the real
dimensions of such vast slides.
Evolution
Profile BAC E15 (Figures 6A-6F) shows facies changes and the
evolution of sequence II at the foot of the slope toward the basin.
In part, BAC E15-1 (Figures 6A and 6D), sequence II, shows the same
characteristics as those described on profile BAC E14-1 (Figure 4A). It
consists of multiple seismic facies units in which the reflections still are
highly organized and separated by many unconformities, generally
indicating slip planes, and affected by many faults.
This part with organized reflections develops very rapidly toward a
chaotic seismic facies unit, where only a few rare, organized reflections can
be discerned (see BAC E15-2, Figure 6B and 6E, shotpoints 700 to 900).
This facies can be seen on all the west-east sections.
On profile BAC E15-2 (Figures 6B and 6E), from shotpoints 900 to the
end, or on profile BAC E14-1 (Figure 4A) toward shotpoint 100, this

chaotic unit intertongues with seismic facies characterized by very


continuous reflections.
The change in the seismic facies from upstream to downstream also is
described on profile CF 110 (Ravenne et al., 1988) for siliciclastic deposits.
This change is explained by the stripping and progressive fluidization of
the slide mass. It is first reflected by a slightly deformed but very faulted
zone with shelf and slope fragments, followed by a very crushed zone with
chaotic reflections, and finally by a turbiditic sedimentation zone where the
units again become organized.
The multiplicity of slide episodes, visible at the foot of the slope (but
not in the chaotic zone), again becomes clear where the chaotic and
organized seismic facies units are intertongued, with each chaotic tongue
representing a distinct episode.

DESCRIPTION OF SEQUENCE III (TURBIDITES AND


CONTOURITES): CHANGES IN ITS SEISMIC-FACIES
UNIT COMPOSITION
Quaternary sequence III reflects the deposition of materials from the
shelf; these materials arrived via the canyon-channel system (Great Bahama
Canyon-Outer Channel of Great Bahama Canyon), as described from the
Indus area (see Ravenne, Coumes, and Esteve, 1988). Thus, complex
turbidity currents are involved, which together with currents sweeping along
the Bahama Scarp establish a contourite system. Morphologically, the
channel-levee system is not symmetrical. A main channel runs along the
Bahama Scarp from the outlet of the canyon. This channel is only bordered
by one levee along its eastern flank (basin side), whose thickness decreases
from the proximal part toward the distal parts, simultaneously with the
increase in its width.
The changes in sequence III from the canyon outlet toward the north
display a wide variety of seismic facies units with such wide differences
that it was difficult to correlate the sequences and the subsequences from
one profile to the next, without a narrow mesh.
On outcrops, it is just as difficult to correlate the same sequencesedimentary objects exhibiting such wide thickness differences. Resolution
power of stratigraphic surveys is inadequate. Moreover, these objects

Deep clastic carbonate deposits of the Bahamas

104

generally are highly reworked and unsuitable for faunal studies.


The change from the south toward the north of sequence III shows the
following main features. On profile BAC E9 (Figures 7A and 7B), between
shotpoints 200 and 400, the channel itself does not contain sediments
belonging to sequence III. The overflow deposit at its eastern side displays a
relief of more than 500 m and the seismic facies units mainly display
continuous reflections. These reflections generally are not wavy except at
the overflow deposit-channel contact, where they also are affected by many
slip faults toward the channel (comparable to those seen on the Indus
levees) (Ravenne, Coumes, and Esteve, 1988).
Farther north, profile BAC E15 (Figure 6) shows a decrease in the level
difference between the channel bed and the uppermost point of the levee.
The channel was initially filled by chaotic seismic facies units (subsequence
III a), and then by seismic facies units with organized reflections of low
continuity. In the levee, the seismic facies is characterized by parallel
continuous reflections, becoming slightly wavy at the approach of the
channel.
Profile BAC E14 (Figure 4) shows that the decrease in the level
differences between the maximum high point of the levee and the channel
bed continues northward and is accompanied by a decrease in the slope. In
the channel, the reflections become better organized. The levee widens, but
it mainly displays a substantial change in seismic facies units, in
comparison with profiles BAC E9 and BAC E15 (Figures 7 and 6). It now
consists of a stacking of facies units of migrant wave hillocks, with
continuous and well-organized reflections.
On profiles BAC E12 and BAC E5 (Figure 3 and 8A-D) there are
practically no level differences between the channel and the levee. The
channel fill and the levee consist of the same seismic facies unit, with
relatively continuous and parallel reflections.
Quick analysis of these two sequences result from material remobilized
by gravity processes but with different masses, transports, and fluidizations,
reveals a wide variety of seismic facies units organized laterally from the
influx zones toward the distal zones. Sequence II, resulting from major
mass slidings, displays a wide chaotic seismic facies unit which is not
reorganized within 20 km of the foot of the slopeafter the material has
been suspended and transported. Only this part can be compared with the
distal deposits of sequence III.
The reader also should compare changes of sequence III in profile BAC
E to of the upper sequence in the Cap Ferret profiles (Ravenne et al., 1988).
Figure 2. Extract of simplified geological map of the western Alps (after Gidon and Richard).
Figure 1. Index map showing bathymetry near the Great Bahama Canyon and shot
point locations for the seismic lines discussed in this chapter. Bathymetry contour
interval in meters.

C. Ravenne, P. LeQuellec, P. Valery and R. Vially

Deep clastic carbonate deposits of the Bahamas

105

COMPARISON WITH MESOZOIC OUTCROPS OF THE


VERCORS OF THE VOCONTIAN TROUGH
Field surveys conducted in southeastern France (Figures 2 and 10) at the
Vercors-Vocontian boundary and in the Vocontian area, show sedimentary
bodies with geometric characteristics (depression fill, thickness variations)
that are closely comparable to those of the sequences of profiles BAC E.
The depositional environment is that of a basin with a northern margin
analogous to that of the seismic profiles analyzed. The structural variations
within the bodies observed show the extent of the redeposition processes,
which sometimes involve mass slides. The sedimentary bodies identified in
these outcrops are of a plurihectometric dimension that is comparable to our
seismic profiles. These sedimentary processes occurred throughout the
Mesozoic, and it appears that a certain permanence prevailed in the location
of the canyons and the main channels.
The different types of deposits resulting from the reworking processes
are shown in chronological order. Photos 1 through 5 illustrate the evolution
of a Late Jurassic canyon and deposits of the supra-fan type (wide channels,
deposition lobes). Photos 6 through 10 illustrate the southward progradation
of the Urgonian shelf. They emphasize the succession of slopes. The contact
of one of the parts with the deep deposits is shown in Photo 11. Photo 12
illustrates the collapse processes that affected the boundary of the Urgonian
shelf. Photo 13 shows the deposits from an Early Cretaceous Canyon fill.
Photos 14 and 15 show the proximal slope base deposits of Late Cretaceous
age. Photo 16 illustrates an onlap contact of carbonate turbidites.
Late JurassicEvolution of Deep Deposits

Figure 10. Simplified geologic map of western Alps, in southeastern France.

C. Ravenne, P. LeQuellec, P. Valery and R. Vially

Photo 1 shows the late Jurassic canyon fill, or rather a major channel
fill, located at the outlet of a canyon. Only the lower part is clearly visible in
the photo, and the maximum thickness exceeds 50 m. It shows a
superposition of increasingly thin lenses toward the top that are increasingly
wide and formed of increasingly fine-grained carbonate material. The
contact occurs by onlap at the edges. The fill lies on a bed about 1 m thick
that is very deformed. This suggests that the canyon masses continued to
slide after their deposition, causing deformation of this bed and,
consequently, the depositional surface became inclined.
The boundaries of the fill lenses are affected by many curved
readjustment faults toward the center of the canyon. For us, this fill and its
characteristics are comparable to the channel fill (subsequences IIIb and
IIIc) of profile BAC E15 (Figure 6), shotpoints 500 to 700. This channel has
a north-south direction and is found at the point marked 2 on the map north
of Gresse en Vercors (Figure 10).
Photo 2 shows that, in the lengthwise direction of the channel, the fill is
still formed by a succession of lenses whose vertical succession is the same
as that of Comboire (Photo 1). The channel continues southward and

widens. The lateral slopes become increasingly more gentle. This


development is comparable to that observed on profiles BAC E.
In the Vocontian area, sedimentary structures are mainly located near the
channels. Among these remarkable structures are the mountains of Ceuse,
Pierre Ecrite, and St. Genis.
Beaudoin (1977) carried out the detailed sedimentological analysis of
the structures of the Vocontian area. This work revealed the reworking that
affected the sediments and identified fossil canyons.
Photo 3 shows characteristic outcrops. A complex asymmetric bar is
observed at the top, with its maximum thickness in the axis of the main
channel. Several lenses coalesce, the thickest and narrowest being located
near the axis and at the bottom. The lenses progressively gain in extension
while generally decreasing in thickness. This photo and photos 4 and 5 are
fairly representative of the channel fill as observed in subsequence IIIb on
profile BAC E14 (Figure 4) between shotpoints 400 and 750, with closely
comparable sizes.
Photo 4 illustrates the more massive appearance of the basal lenses and
still shows the asymmetric aspect of the outcrop from the channel axis
toward the edges.
Photo 5 is situated slightly east of Photo 4 and shows a water escape
structure comparable to the one described by Labaume (1983) in the
Spanish Pyrenees. These structures combine with detailed sedimentological
analyses to suggest that the canyon fill results from mass slides. The water
escape forms suggest the trapping of the liquid tongue at the lower part of
the canyon (see Ravenne and Beghin, 1983).
Urgonian-Shelf Progradation
Photos 6 to 10 illustrate the southward progradation of the Urgonian
shelf. Detailed stratigraphic and sedimentological analyses formed the
subject of the outstanding work of Arnaud (1981) and Arnaud-Vanneau
(1980), which helped to identify southward progradation of the shelf.
A vertical sequence of the outcrops south of the Vercors region (Photos
6 to 8) shows successively very steeply sloping limestone beds at the base
that are separated by marly layers. These steeply sloping layers depict a
paleoslope overlain by increasingly more gently sloping, prograding beds
with increasingly thinner marly intercalations. These prograding beds are
covered by horizontal or slightly sloping limestone banks representing the
innermost parts of the shelf. Their sloped termination is found further south.
Photos 8 and 10 illustrate the dip variations in the foresets. The
progradation banks form first-order foresets on a large scale, but they
themselves consist of more inclined, second-order foresets.
The size of these prograding beds is perfectly compatible with the
seismic evidence. Such foresets are observed all around the southern part of
the Vercors region and the present topography is very similar to the

Deep clastic carbonate deposits of the Bahamas

131

paleotopography. Thus, the mountains located on either side of the Archiane


cirque (point 5)Mount Glandasse (point 4), and Mount Benevise (point 7;
see Figure 10)are headlands from which the progradations trend
predominantly southward, but also laterally, on either side of the headlands.
UrgonianPassage to Deep Deposits, and Stripping Processes
Affecting the Shelf
Photo 11 shows the termination (point 7, see Figure 10) of the mountain
bordering the Archiane cirque (point 5, Figure 10) to the east. The bodies
observed are still at a scale compatible with the seismic sections. The
margin of a paleoshelf (about 100 m thick) is observed at A, whose upper
slant is extended by paleoslope B. At the foot of paleoslope B, deep-sea-fan
type deposits are found, located at least 400 m below the paleolevel of the
water (taking the shelf surface A as the minimum reference). This unit is
covered by marly limestones on which a new shelf unit, D, lies. Worth
noting is the further thickening of this unit near the end of the shelf. The
presence of small listric faults in unit D relates to the differential
compaction of the marly limestones, to allow some extra thickening.
The slope transition is no longer observed. Subaerial erosion was active
and it is probable that the final advance of the shelf could not be stabilized.
This image seems fairly similar (but reduced) to that of the Bahama Scarp
during the supply of sequence II with the slides of the slope and shelf
deposits. It also appears that this process can be associated with the deposits
observed at the southern end of the Vercors region (Photo 12), where the
Urgonian shelf fragments appear to emphasize collapses on a deformed bed
of Hauterivian marly limestones.
Photo 13 shows Mont Aiguille (point 3 on Figure 10). This is a large
limestone mass lying abruptly on the underlying marly limestones. There is
no transition, whereas below the shelf with horizontal deposits, the
transition is marked by increasingly inclined prograding foresets,
intercalated with marly limestones (Photo 7). The internal structure also
shows several lenses. The configuration is comparable to that of Photo 2.
Also involved here is a main channel or even canyon fill, which could be
compared to the main channel fill of BAC E15 (Figure 6) in sequence III
(shotpoints 600 to 700).
Upper CretaceousSlope Base Deposits
Photos 14 and 15 illustrate relatively proximal slope base deposits,
closely comparable in their situation to the deposits of sequence II when
they display a chaotic seismic facies. Here the lenses contain boulders of
multidecimetric size (Photo 15; Gats conglomerates, point 8 on Figure 10).
The outcrop does not permit observation of longitudinal change of the
facies from north to south. However, the vertical and lateral changes (center

C. Ravenne, P. LeQuellec, P. Valery and R. Vially

to east) reveal a very progressive change of these conglomerateswhich fill


a depression about 2 km wide at the baseto a unit formed of much wider
and thinner lenses, consisting of less coarse material, and finally to
turbiditic type deposits. The complete change occurs over a thickness of
about 700 m and shows a passage from relatively chaotic deposits to very
well-organized deposits. The depositional processes, the environment, and
the evolution are very similar to those observed on profiles BAC E15, E14,
E12, and E5 (Figures 6, 4, 3, and 8, respectively), in sequence II, at the
passage between the chaotic facies units and the facies units with wellorganized reflections, going toward the basin.
Photo 16 shows an onlap contact of limestone gravity deposits, very
similar to those of photo 3, lying in an onlap position on the underlying
marly limestones. This photo shows, in a carbonate environment, a contact
identical to the one observed at Chalufy in the siliciclastic environment
(Cap Ferret, see Ravenne et al., 1988) and illustrates the onlap contacts of
subsequence III on the slope (profiles BAC E12 and BAC E5; Figures 3 and
8).

CONCLUSIONS
The seismic profiles of the foot of the Bahama Scarp help to identify the
scale of the remobilization processes for deposits situated at the foot of the
slopes and in the basins. Two main processes are responsible for the
construction of these deposits. The first is associated with mass slides
affecting the margin of the shelf and the slope. The second is associated
with the currents active in sorting and redeposition of the original material.
The surveys conducted in the Vocontian area and against the southern
edge of the Vercors region show that here similar processes predominate.
The sedimentary bodies display sizes comparable to the bodies identified on
seismic profiles, and are arranged in accordance with dominant processes
occurring near the slope. The analysis of the sedimentary bodies identified
in the outcrop thus allows an easier interpretation of the seismic facies units,
provided their lateral and vertical changes have been identified.
Seismic profiling also reveals sedimentary bodies whose dimensions
generally are very large in comparison with the sedimentary bodies
identified in conventional field surveys. These bodies have therefore only
rarely been investigated and identified. In this shelf margin environment,
seismic profiles also point out the scale of the reworking processes.

ACKNOWLEDGMENTS
We thank the Offshore Petroleum Study Committee (Comite dEtudes
Petrolieres Marines), which allowed the seismic survey, the crew of NO
Resolution, and the geophysicists of IFP who helped to collect and process
the data.
Our special thanks go to Compagnie Francaise des Petroles et Institut
Francais du Petrole for their aid and support provided during this study.
We also thank F. Euriat and J. Rabate for their active participation in the
analysis of the BACAR profiles.

REFERENCES CITED
Arnaud, H., 1981, De la plateforme urgonienne au bassin vocontien: le BarremoBedoulien des Alpes occidentales entre Isere et Buech: Geologie Alpine,
Memoir 12, 3 vols. 804 p.
Arnaud-Vanneau,A.,1980,Micropaleontologie, paleocologie et sedimentologie dune
plateforme carbonatee de la marge passive de la Tethys:lUrgonien du Vercors
septentrional et de la Chartreuse:Geologie Alpine, Memoir 11, 3 vols., 873 p.
Beaudoin, B., 1977, Methodes danalyse sedimentaire et reconstitution du bassin:
le Jurassique terminal-Berriasien des chaines subalpines meridionales:
Ph.D. thesis, Cannes, 339 p.
Johns, D.R., E. Mutti, J. Rosell, and M. Seguret, 1981, Origin of a thick,
redeposited carbonate bed, in Eocene Turbidites of Hecho Group, SouthCentral Pyrenees, Spain: Geology, v. 9, p. 161-164.
Labaume, P., 1983, Evolution tectono-sedimentaire et mega-turbidites du bassin
turbiditique eocene sud-pyreneen entre les transversales Somport-Jaca et le
Pic dOhry Sierra de Leyre: These de 3eme cycle, U.S.T.C.,
Montpellier.
Labaume, P., E Mutti, M. Seguret, and J. Rosell, 1983, Megaturbidites
carbonatees du bassin turbiditique de leocene inferieur et moyen sudpyreneen: Bull. Soc. Geol. France (7), t. XXV, no. 6, p. 927-941.
Ravenne, C., and P. Beghin, 1983, Apport des experiences en canal a
linterpretation sedimentologique des depots de cones aquatiques
sousmarins: Rev. IFP, 38.3, p. 279-297.
Ravenne, C., P. LeQuellec, and P. Valery, 1985, Depots carbonates profonds des
Bahamas, in A. Mascle, ed., Geodynamique des Caraibes (Paris): Editions
Technip, Paris.
Ravenne, C., M. Cremer, P. Orsolini, and P. Riche, 1988, Mass slides and turbidite
type deposits recognized by offshore seismic prospecting: Cap Ferret
depression and at the outcrop: Gres dAnnot Series, this volume.
Ravenne, C., F. Coumes, and J.P. Esteve, 1988, Influences of relative variations of
sea level on depositional modes of the shelf and deep-sea fan of the Indus:
this volume.
Ravenne, C., P. LeQuellec, and W. Schlager, In preparation, Seismic stratigraphy
of two carbonate fan areas east of the Bahamas.
Vail, P.R., R.M. Mitchum, et al, 1977, Seismic stratigraphy and global changes of
sea level, in C.E. Payton, ed., Seismic stratigraphyapplications to
hydrocarbon exploration: AAPG Memoir 26, p. 40-212.

GEOPHYSICAL PARAMETERS
FOR SEISMIC PROFILES
Energy Source:
Stacking Multiplicity:
Number of channels recorded:
Interval between Input Channels:
Minimum offset Distance:
Maximum offset Distance:
Frequency Filtering:
Migration:

Hydraulic Flexichoc
24
4B
25 m
260 m
1435 m
20 - 60
Yes

Deep clastic carbonate deposits of the Bahamas

132

ERODED CARBONATE PLATFORM MARGIN


THE BLAKE ESCARPMENT OFF SOUTHEASTERN UNITED STATES

WILLIAM P. DILLON,
ANNE M.TREHU,
PAGE C. VALENTINE
and
MAHLON M. BALL
U.S. Geological Survey
Woods Hole, Massachusetts
INTRODUCTION
Steep carbonate escarpments are common on the sea floor and in the
rock record. They are created by construction of carbonate platforms and by
subsequent erosion that oversteepens their faces. The steep slopes, nearly
vertical in places, produce significant problems in processing and
interpreting common-depth-point (CDP) multichannel seismic-reflection
profiles, because the normal assumption of continuous, flat-lying layers is
not fulfilled. Carbonate platforms tend to be difficult areas for profiling
anyway, because of rough surfaces and cavities produced by karstification,
and occurrence of common high-velocity layers shallow in the sedimentary section.
The Blake Escarpment is a well-developed, steep carbonate escarpment
on the sea floor east of Florida on the U.S. eastern continental margin
(Figure 1). The U.S. Geological Survey (USGS) obtained several
multichannel seismic reflection profiles across this escarpment as part of its
program to survey the continental margin; part of one of these profiles,
profile TD4, is presented here. The regional location of this section is shown
in Figure 1 and location and detailed bathymetry are shown in Figure 2.
In addition to multichannel and single-channel seismic profiling, we
made three dives with the research submersible ALVIN to observe and
sample the escarpment along profile TD4. Dive transects of the escarpment
also were made along two other profiles (sites A and C, Figure 1). These
observations and analyses of rocks from the escarpment have been valuable
in interpreting seismic profiles. Other discussions of multichannel seismic
profiles in this region already have been presented (Dillon et al., 1976,
1979a, 1979b, 1985; Dillon and Paull, 1978; Shipley et al., 1978; Buffler et
al., 1978; Schlee et al., 1979; Sheridan et al., 1979, 1981).

W.P. Dillon, A.M. Trehu and P.C. Valentine

GEOLOGIC SETTING
The morphology of the continental margin off the southeastern United
States consists of a normal continental shelf and upper slope, but the slope
is interrupted by the broad, flat Blake Plateau (Figure 1). The plateau is
underlain by one of the four major basins of the eastern U.S. continental
margin, the Blake Plateau basin. The basin contains a section of
sedimentary rocks as thick as 13 km, much of which are considered to have
accumulated in a carbonate platform setting (Dillon and Popenoe, 1986).
We believe that shallow-water carbonate platform deposition ceased at the
end of the Early Cretaceous, and deep-sea erosion, which served to steepen
the escarpment, probably has been most effective from the Tertiary to the
present.

DATA COLLECTION AND PROCESSING


Seismic profile TD4 (Figure 7) was collected and processed by
Teledyne Exploration Company under contract to the USGS. (Any use of
trade names and trademarks in this report is for descriptive purposes only,
and does not constitute endorsement by the U.S. Geological Survey.) The
seismic source consisted of four 540-cu-in (8.85-liter) air guns fired every
50 m. The streamer was 3600 m long and contained 48 active sections; the
24 sections nearest the ship were each 50 m long and the 24 farthest from
the ship were each 100 m long. In processing, the data were demultiplexed,
and parameter selections (such as filter tests, scaling tests, etc.) were
completed. Then basic processing was applied, including binary gain
recovery, spherical divergence correction, predictive deconvolution, velocity
analysis at 3-km intervals, normal-moveout correction, 48-fold commondepth-point stack, poststack time-variant deconvolution, and time-variant
filtering and scaling. This processing resulted in the records of Figures 7
and 8.

INTERPRETATION OF PROFILE
The most notable feature of seismic profiles in this area is the Blake
Escarpment (Figures 7 and 8). This is a very steep cliff, although it must be
Figure 1. Locations of profile TD4 section, submersible dive sites (sites A, B, and C)
and bathymetry of continental margin of the southeastern United States.

kept in mind that the vertical exaggeration in the profile is 4:1 at the sea
floor. (Vertical exaggerations in time sections become progressively smaller
beneath the sea floor as velocities increase.)
A series of dives to observe and sample the escarpment were made
aboard the deep-diving submersible ALVIN. At the location of profile TD4,
three dives were used to make a continuous transect from 4000 m up to
2617 m. The observations show that the sea floor is a steep, stepped slope,
consisting of eroded outcrops of essentially flat-lying limestone strata with
intervening areas of talus and biogenic sand. Figure 5 shows the escarpment
sketched from dive observations at an approximate vertical exaggeration of 2:1.
A typical outcrop is shown in Figure 6; this photograph emphasizes the
importance of jointing in the morphology of the escarpment. Jointing is
very common, and we believe that it probably is the result of uneven stress
release on the rocks, caused by removal of material to seaward by erosion.
The limestones sampled represent shallow water, bank interior deposits
(M. Arthur, E. Shinn, personal communication, 1981). Ages are based on
calcareous nannofossils and are shown on the seismic section (Figure 8) and
in a summary diagram (Figure 4), which includes the more extensive dating
afforded by samples from other dive transects and Deep Sea Drilling Project sites.
The strata are of Early Cretaceous age, down to 4000 m, the greatest
depth of sampling by ALVIN. Insofar as the escarpment exposes rocks to a
depth of nearly 5000 m, and the dated samples at 4000 m represent rocks
that are nearly the oldest of the Cretaceous, it seems probable that Jurassic
rocks may be exposed at the base of the escarpment.
At the foot of the Blake Escarpment and buried by the sediments of the
Blake-Bahama basin, lies a bench-like feature (Figures 7 and 8) that was
mapped semi-continuously along the base of the escarpment, and that is
considered to be a remnant left after erosion and retreat of the seaward part
of the plateau (Paull and Dillon, 1980). Although this view was
controversial (Sheridan, 1981; Paull and Dillon, 1981), further studies
appear to support an erosional retreat model for the Blake Escarpment
(Dillon and Popenoe, 1986).
West (left) of the Blake Escarpment, a series of steep, downward
flattening faults are interpreted (Figure 8). These imply extension, with
minor slumping. They may be in part a response to the removal by erosion
of rock at thedifferent from other published information (and from Figure

Eroded carbonate platform marginthe Blake escarpment

140

Figure 5. Sketch of a section of the Blake Escarpment at profile TD4, based on


observations from submersible ALVIN. See Figure 8 for location of dive observations
relative to seismic profile.
Figure 6. Photograph of escarpment at 3892 m at site C, showing jointed stepped nature of the cliff.

Figure 4. Ages of rocks sampled on the Blake Escarpment from ALVIN (Sites A, B,
and C) and stratigraphic information from drill sites on the Blake Spur (see Figure 1
for locations).

foot of the escarpment and a resultant change in internal stress, as we also


suggested for the jointing.
A well-defined reef structure is apparent in the profile beneath the outer
Blake Plateau at shotpoints 8500 to 8960. The main reef structure outlined
in the interpretation (Figure 8) is a region of suppressed reflections and lack
of coherent continuous reflections. We correlate this reef structure to reefs
crossed by several other profiles in similar locations on the outer Blake
Plateau. South of profile TD4, at dive site C (Figure 1), the erosion of the

W.P. Dillon, A.M. Trehu and P.C. Valentine

escarpment truncated a reef of this system, and rudists of Albian-Aptian age


are present (Figure 4; N.F. Sohl, written communication, 1981).
We conclude that the structure observed on TD4 is a rudist-containing
reef as well. The reef was situated at the edge of the shallow carbonate
platform in Albian-Aptian time and represents a step-back of the platform
edge from an earlier position, as indicated by the structure displayed in
profile TD4 (Dillon et al., 1985; Dillon and Popenoe, 1986). A small
unwarping of reflections just west of the escarpment at about 3.8 to 4.3 sec
depth (shot-points 9201 to 9241) may be the remnant of a larger reef that
existed to the eastward and has been eroded away.
The postrift unconformity (PRU, Figure 8) is the deepest mappable
surface that crosses the continental margin, and is identified on the basis of
the seismic profile alone. The closest site at which the PRU has been drilled
is on the continental shelf, more than 300 km away. The PRU is considered
to be the top of synrift sedimentary rocks, rift-related volcanic rocks,
Paleozoic metamorphic rocks and crystalline basement, all truncated by
erosion before the continental margin subsided below sea level.

MODELING THE EFFECT OF THE BLAKE


ESCARPMENT ON SEISMIC DATA
To illustrate the effect on seismic data of the large lateral velocity
variation resulting from the presence of the Blake Escarpment, synthetic
data were generated using the AIMS seismic modeling program (Advanced
Interpretive Modeling Systema trademark of Geoquest International Inc.).
Figure 3A shows the model for which the seismograms were calculated.
Stars indicate points that were defined to be diffraction sources; large
numbers indicate the interval velocity in each block; and small numbers are
the layer numbers. The topography of the model was simplified from the
topography along line TD4, taken from the seismic profile and dive
observations. The assumed velocity structure was based on large-offset
seismic data collected on the Blake Plateau using air guns and ocean-bottom
seismometers (Trehu, 1984). The layering of high and low velocity zones

shown in the model is supported by the correlation between the large-offset


seismic data, CDP seismic-reflection data, and observations of the Blake
Escarpment made from ALVIN. The submersible observations suggest that
the high velocity zones correspond to layers of massive lithified limestone
and dolomite, whereas the low velocity zones correspond to less
consolidated deposits.
Figure 3B shows normal incidence raypaths for this model, and Figure
3C shows a synthetic single-channel seismic record. The seismic image of
the escarpment is severely distorted because of diffractions from the face of
the escarpment. Note that no background noise, multiples (water-bottom or
peg-leg), or P-to-S-to-P converted arrivals are included in this simple
model; these are factors that also contribute to degrading the seismic data
along TD4.
The effects of lateral heterogeneity are even more serious for
multichannel CDP data and can degrade the data further, compared to

Eroded carbonate platform marginthe Blake escarpment

142

single-channel data. Figure 3D shows raypaths for CDP gathers at several


positions along the escarpment and Figure 3E shows the corresponding
seismograms. Rays were traced to simulate a 12-channel streamer with an
initial offset of 150 m, 322-m spacing between channels, and a total offset
of 3692 m. This corresponds to approximately the same total offset and
25% of the data density of line TD4.
CDP I is more than one streamer length from the upper edge of the
escarpment; therefore the effect of the topography is not yet felt. The
interlayering of low and high velocity zones, however, causes the first
arrival hyperbolae from various layers to cross, which will distort and smear
the data when normal moveout and stacking are applied.
At CDP II, near the edge of the escarpment, we see shadow zones at
certain ranges, and the travel times depart from the hyperbola assumed
when stacking CDP data, resulting in additional smearing of the image.
At CDP III, over the escarpment, the data are severely distorted; and at
CDP IV, over the abyssal plain, hyperbolic returns from irregularities on the
escarpment interfere with the arrivals from below. Vertical exaggeration in
Figures 3A-3D is approximately one. To facilitate comparison with the data,
Figures 3F and 3G show the model and synthetic single-channel section
with a vertical exaggeration of four.
This modeling indicates that zones of apparently disturbed reflections
will be observed in regions of pronounced lateral velocity variations. Such
distortion is inherently produced by multichannel processing, when the
assumptions of flat-lying, continuous reflectors are not fulfilled.

IMPLICATIONS OF MODELING AND FURTHER


PROCESSING TO INTERPRETATION
In unmigrated profiles (Figures 7 and 8), the face of the Blake
Escarpment is obscured by hyperbolic diffractions that are generated at
irregularities on the slope, as modeled in Figure 3C. The irregularities, in
part, are those depicted in Figure 5. Migration of the data (Figure 9) shows
the face of the escarpment by causing the collapse of the diffractions back
to their source.
The diffractions not only obscure the cliff face, they also extend to the
left and disrupt reflections from strata within the platform, as Figure 3C
predicts. Diffractions from the face of the upper step of the escarpment
(shotpoints 9050 to 9120) also interfere with reflections from strata in the
region west of the main part of the escarpment below (about) 3 sec.
Migration (Figure 9) reduces some of the disruption and clarifies structure
just behind the escarpment by removing the interfering hyperbolic
diffraction returns.
Disruption and suppression of reflections within reefs in profiles is
common, as noted above for the Albian-Aptian reef as shotpoints 8500 to
8960 (Figure 8). The disruption and weakening of reflections in the region

W.P. Dillon, A.M. Trehu and P.C. Valentine

behind the Blake Escarpment have led to the interpretation of reef structures
there. However, the westward extension of diffractions as modeled in Figure
3C, and impact of complex raypaths on the processed CDP seismic record
as indicated in Figure 3D, probably account for the effect. Except where
other structural evidence for reefs existssuch as at depths of 3.8 to 4.3 sec
just behind the escarpmentwe need to be careful when we interpret reefs
adjacent to steep slopes.
In addition to the migrated time section, a migrated depth section is
shown (Figure 10). Unfortunately, the depth-converted profile displays
some significant errors in imaging the deeper structure. These errors
presumably are due to the abrupt lateral change in velocity produced by the
Blake Escarpment. Whether the errors in chosen velocities are due to
inappropriate averaging or to problems with velocity analyses resulting
from the complex raypaths (Figures 3D and 3E) is not known. Near the
steep lower part of the escarpment (shotpoints 9100 to 9250), the chosen
velocities seem to be too low, with error increasing toward the escarpment.
This results in the apparent pulling up of deeper reflections into a
(probably) nonexistent monoclinal structure in the region of 4 to 8 km
depth.
Just seaward of the escarpment (shotpoints 9300 to 9350), strata and the
sea floor appear to dip toward the escarpment, giving the appearance of a
moat next to the foot of the cliff. The moat does not appear in the time
sections, indicating that velocities used for depth conversions are too high in
that area, even for the water column.
Away from the escarpment, the depth-converted section is more useful.
It points out that, although the PRU appears in time sections to be shallower
than the adjacent Blake-Bahama basin floor, it truly is much deeper (Figure
10). The dip of reflections seaward of the reef, beneath the step of the inner
Blake Escarpment (shotpoints, or SP, 9000 to 9200), is almost completely
due to increased water depth. As shown by the depth-converted section
(Figure 10), the strata are approximately horizontal in this region. This
affect is well modeled in Figure 3 (compare 3B to 3C).

CONCLUSIONS
The Blake Escarpment is a steep, underwater cliff in Lower Cretaceous
rocks on the seaward edge of the Blake Plateau carbonate platform. Seismic
profiles show that the escarpments structure is consistent with a history of
construction followed by erosion and retreat. Steep carbonate escarpments
are common on the sea floor and cause major problems in the processing
and interpretation of seismic profiles. At such locations the usual
assumption of flat-lying, continuous velocity units is incorrect and the
abrupt lateral change in velocity structure at the cliff face seriously degrades
multichannel CDP seismic profiles.
For energy that penetrates the sea floor, the steep cliff produces a

variable pattern of refraction, and for energy that is reflected from the sea
bottom, the irregular slope creates overlapping and interfering hyperbolic
diffractions. These phenomena produce disturbed or weakened reflections
adjacent to the cliff, which might be interpreted mistakenly as reefs or
tectonically disturbed areas. The problems of determining correct velocity
structure near a cliff face also create major difficulties in creating depthconverted sections.

ACKNOWLEDGMENTS
This paper benefited greatly from reviews by Deborah Hutchinson and
Kim Klitgord. We thank Margaret Clare Wengler, Elizabeth Winget, Patricia
Forrestel, Jeff Zwinakis, and Dann Blackwood for their assistance.

REFERENCES CITED
Buffler, R.T., T.H. Shipley, and J.S. Watkins, 1978, Blake continental margin
seismic section: American Association of Petroleum Geologists, Seismic
Section No. 2. 1 sheet.
Dillon, W.P, and C.K. Paull, 1978, Interpretation of multichannel seismicreflection profiles of the Atlantic continental margin of the coasts of South
Carolina and Georgia: U.S. Geological Survey Miscellaneous Field
Investigations Map MF-936.
Dillon, W.P, C.K. Paull, R.T. Buffler; and J.P. Fail, 1979a, Structure and
development of the Southeast Georgia Embayment and northern Blake
Plateau: Preliminary analyses, in J.S. Watkins, L. Montadert, and P.W.
Dickerson, eds., Geology and geophysical investigations of continental
margins: American Association of Petroleum Geologists Memoir 29, p. 2741.
Dillon, W.P, C.W. Poag, P.C. Valentine, and C.K. Paull, 1979b, Structure,
biostratigraphy, and seismic stratigraphy along a common-depth-point
seismic profile through three drill sites on the continental margin off
Jacksonville, Florida: U.S. Geological Survey Miscellaneous Field Studies
Map MF-1090.
Dillon, W.P. and Peter Popenoe, 1986, Development of the continental margin of
the southeastern United StatesThe Blake Plateau basin and Carolina
Trough in R.E. Sheridan, and J.A. Grow, eds., The geology of North
America: the Atlantic continental margin, U.S.: Geological Society of
America, Geology of North America, v. I-2, in press.
Dillon, W.P, R.E. Sheridan, and J.P. Fail, 1976, Structure of the western Blake
Bahama basin as shown by 24 channel CDP profiling: Geology, v. 4, p.
459-462.
Dillon, W.P, C.K. Paull, and L.E. Gilbert, 1985, History of the Atlantic continental
margin off Florida: the Blake Plateau basin; in C.W. Poag, ed., Geologic
evolution of the United States Atlantic margin: New York, Van Nostrand
Reinhold, p. 189-215.
Gilbert, L.E., and W.P. Dillon, 1981, Bathymetric map of the Blake Escarpment:
U.S. Geological Survey Miscellaneous Field Investigations Map MF-1362.
Paull, C.K., and W.P. Dillon, 1980, Erosional origin of the Blake Escarpment: An

alternative hypothesis: Geology v. 8, p. 538-542.


Paull, C.K., and W.P. Dillon, 1981, Reply to Comment on erosional origin of the
Blake Escarpment: An alternative hypothesis: Geology, v. 9, p. 339-341.
Schlee, J., W.P. Dillon, and J.A. Grow, 1979, Structure of the continental slope off
the eastern United States, in L.J. Doyle, and O.H. Pilkey, eds., Geology of
continental slopes: Society of Economic Paleontologists and Mineralogists
Special Publication 27, p. 95-117.
Sheridan, R.E., 1981, Comment on erosional origin of Blake Escarpment: An
alternative hypothesis: Geology, v. 9, p. 338-339.
Sheridan, R.E., J.I. Windisch, J.I. Ewing, and P.L. Stoffa,
1979, Structure and stratigraphy of the Blake Escarpment based on seismic
reflection profiles, in J.S. Watkins, L. Montadert, and P.W. Dickerson, eds.,
Geological and geophysical
investigations of continental margins: American Association of Petroleum
Geologists Memoir 29, p. 177-186.
Sheridan, R.E., J.T. Crosby, G.M. Bryan, and P.L. Stoffa, 1981, Stratigraphy and
structure of the southern Blake Plateau, northern Florida Straits and
northern Bahama platform from multichannel seismic-reflection data:
American Association of Petroleum Geologists Bulletin, v. 65, n. 12, p.
2571-2593.
Shipley, T.H., R.T. Buffler; and J.S. Watkins, 1978, Seismic stratigraphy and
geologic history of the Blake Plateau and adjacent western Atlantic
continental margin: American Association of Petroleum Geologists
Bulletin, v. 62, p. 792-812.
Trehu, A.M., 1984, Effects of bottom currents and uppermost sediment thickness
on ocean-bottom seismometer data (abs.): EOS, Transactions American
Geophysical Union, v. 65, p. 1014.

Eroded carbonate platform marginthe Blake escarpment

143

EROSION OF THE SOUTHERN FLORIDA ESCARPMENT1

WILLIAM CORSO,
RICHARD T. BUFFLER,
and
JAMES A. AUSTIN, JR.
University of Texas at Austin
Institute for Geophysics
Austin, Texas
INTRODUCTION
Geophysical and geological studies of the Early Cretaceous carbonate
platform margin exposed along the Florida escarpment show that its steep
relief is the result of complex tectonic, eustatic, depositional and erosional
processes. The morphologies of both the escarpment and platform margin
change dramatically from northwest to southeast (Bryant et al., 1969; Corso
and Buffler, 1985; Twichell et al., 1986). The escarpments relief increases
to the south from about 1000 m to more than 1500 m between 2830N and
2340N. South of 27N, canyons incise the escarpment about every 7 to 15
km (Figure 1; Twichell et al., 1986). North of 2830N, the platform margin
is rimmed and has an accretionary slope that is buried beneath Upper
Cretaceous to Holocene deep-water sediments (Figure 1; Addy and Buffler,
1984). A sediment bypass slope developed, and is exposed along
the escarpment from 2830N to 2340N before becoming buried
beneath the western Straits of Florida (Figure 1; Locker and Buffler, 1983;
Phair, 1984; Corso and Buffler, 1985).
Several observations along the Florida escarpment south of 27N,
including the recovery of outcropping platform interior facies, the lack of a
seismically identifiable platform margin facies, and the occurrence of
canyons and possible slumps near its base, all suggest that this part of the
escarpment was eroded by as much as 5 to 10 km (Bryant et al., 1969;
Antoine et al., 1974; Freeman-Lynde, 1983; Locker and Buffler, 1983). All
of these observations, though, provide indirect measurements of the amount
of erosion.

Figure 1. Map of the eastern Gulf of Mexico, showing locations of seismic lines,
dredges (Freeman-Lynde, 1983), ALVIN dives (Paull et al., 1984), and DSDP sites 535
and 540 (Buffler et al., 1984).

1This is the University of Texas Institute for Geophysics Contribution No. 658.

W. Corso, R.T. Buffler and J.A. Austin, Jr.

Erosion of the southern Florida escarpment

149

Our understanding of possible erosional mechanisms affecting


carbonate escarpments, in general, is poor. It is important, therefore, to
assess accurately how much erosion has occurred along the southern Florida
escarpment.
We define the amount of erosion as the distance between the base of the
estimated Early Cretaceous slope and the base of the present escarpment. In
this paper, we use the techniques of seismic stratigraphy to illustrate direct
measurements of this erosion. First, we present two unmigrated, 24-fold
seismic reflection profiles across uneroded parts of the Early Cretaceous
platform margin (lines AG-4 and SF-9; Figure 2). We then use these profiles
as depositional models to interpret two other profiles, shown both in time
and time-migrated sections, which were collected across eroded parts of the
southern Florida escarpment (lines GT2-11 and GT2-24; Figures 3,4,5, and
6). Rocks collected from the escarpment and Deep Sea Drilling Project
(DSDP) boreholes in the adjacent basin provide geologic information for
this study (Figure 1). Acquisition and processing information for each line
is contained in the Appendix.

SEISMIC STRATIGRAPHY

facies. Chaotic reflections and associated diffraction hyperbolae along the


escarpments edge are interpreted as platform margin facies, whereas flatlying reflections that are bankward of the escarpments edge are interpreted
as platform interior facies (Figure 2; Bryant et al., 1969; Garrison and
Martin, 1973; Mitchum, 1978; Locker and Buffler, 1983; Addy and Buffler,
1984; Corso and Buffler, 1985). The succession of seismic facies patterns
correlates well with generic depositional models for carbonate platform
margins (e. g. Wilson, 1975; Schlager and Ginsburg, 1981; Read, 1985).
In the western Straits of Florida, the Early Cretaceous platform margin
is buried, but its morphology and seismic facies patterns are very similar to
those along the northern Florida escarpment (Figure 2). Below the MCU on
seismic line SF-9, dipping reflections at the base of the platform slope pass
upslope into a narrow zone of chaotic reflections at the platform margin.
Both flat-lying and chaotic reflections occur bankward of the platform
margin. Phair (1984) also interpreted these seismic facies as toe-of-slope,
margin, and platform interior facies (Figure 2).
The occurrence of toe-of-slope facies and their continuity toward the
platform margin, along with apparently in-place platform margin facies,
suggest that little, if any, erosion has affected the Early Cretaceous platform
margin either along the northern Florida escarpment or in the western
Straits of Florida (Bryant et al., 1969; Locker and Buffler, 1983; Phair, 1984).

Generalized Geologic Setting

Southern Florida Escarpment

Previous geophysical studies identified a prominent high-amplitude


reflection that truncates reflections below it near the eastern and southern
margins of the deep Gulf of Mexico basin but becomes conformable in the
deep, central Gulf of Mexico basin (Buffler et al., 1980; Faust, 1984;
Schlager et al., 1984; Shaub et al., 1984). This reflection and unconformity
was termed the MCU (Mid-Cretaceous Unconformity), and is tentatively
correlated with a rapid drop and subsequent rise in sea level during the
middle Cenomanian (Buffler et al., 1980; Faust, 1984; Shaub et al., 1984).
The MCU corresponds with a Late Cretaceous hiatus of 30 m.y. at DSDP
site 540 in the southeastern Gulf of Mexico basin, so its age may range from
early or middle Cenomanian through late Paleocene (Buffler et al., 1984).
The MCU also is recognized on the adjacent margins of the Gulf of Mexico
basin, where it again represents a Late Cretaceous hiatus (Worzel et al.,
1973; Mitchum, 1978; Buffler et al., 1980; Locker and Buffler, 1983; Addy
and Buffler, 1984). The MCU is interpreted as a major stratigraphic turning
point in the geologic history of the Gulf of Mexico, corresponding to the
final drowning of the Early Cretaceous platforms on the west Florida shelf
and Campeche Bank (Schlager et al., 1984).
Seismic line AG-4 illustrates an Early Cretaceous platform margin with
a bypass slope (Figure 2). Locker and Buffler (1983) interpret dipping
reflections below the MCU near the base of the escarpment as toe-of-slope

Based on comparisons with the uneroded margin, we interpret the preMCU sequence of dipping reflections near the base of the southern Florida
escarpment (line GT2-11; Figures 3 and 4) as toe-of-slope facies. Using the
results from DSDP sites 535 and 540 (Figure 1), Phair (1984) also
interpreted a similar sequence on line GT2-24 as Lower Cretaceous toe-ofslope facies (Figures 5 and 6). The pre-MCU toe-of-slope reflections,
however, do not exhibit the same continuity upslope on lines GT2-11 and
GT2-24 that coeval sequences illustrate on lines AG-4 and SF-9 (Figure 2).
These dipping toe-of-slope reflections are truncated by the MCU from the
base of the escarpment for 6.9 km basinward on line GT2-11, compared
with 10.3 km of truncation evident on line GT2-24 (Figures 3-6). The
truncation of pre-MCU toe-of-slope facies is more clearly evident on
seismic line GT2-11 (Figures 3 and 4), but on line GT2-24 the pre-MCU
toe-of-slope reflections pass into a zone of more chaotic reflections near the
base of the escarpment (Figures 5 and 6). Several alternative interpretations
are possible for these chaotic reflections: (1) scattering of acoustic energy
along the irregular surfaces of overlying strata obscures the pre-MCU
reflections; (2) faulting is present within the pre-MCU strata; or (3) preMCU margin facies are present. We prefer the first interpretation, based on
comparisons with the uneroded platform margin (lines AG-4 and SF-9;
Figure 2).

W. Corso, R.T. Buffler and J.A. Austin, Jr.

The recovery of Lower Cretaceous (Aptian-Albian ?) limestones and


dolomites in dredge hauls (Freeman-Lynde, 1983) and ALVIN dives (Paull
et al., 1984) on the southern Florida escarpment near lines GT2-11 and
GT2-24 (Figure 1) further suggests that the Early Cretaceous platform
margin was eroded, exposing platform interior facies (Figures 3-6 show
sampling intervals projected onto the seismic lines). Mudstones,
wackestones, and packstones are the principal types of rocks recovered and
indicate restricted marine depositional environments (Freeman-Lynde,
1983).
Within the platform, pre-MCU flat-lying reflections can be traced from
the platform interior to the escarpments edge between reflection times 2.7
and 3.4 sec on line GT2-11 (Figures 3 and 4). The recovery of restricted
marine facies from this part of the escarpment corroborates our
interpretation that the flat-lying reflections exposed at the escarpments
edge represent truncated platform interior facies (Figures 3 and 4). A narrow
zone of chaotic reflections also is evident at the escarpments edge on line
GT2-11 below 3.4 sec (Figures 3 and 4). These reflections may represent
remnants of a platform margin facies. A minor fraction of the rocks
recovered from the escarpment also consisted of skeletal packstones and
grainstones, suggesting occasional development of more open marine
depositional environments (Freeman-Lynde, 1983).
A broad zone of chaotic reflections is evident within the platform on
line GT2-24 (Figures 5 and 6). The three alternative interpretations stated
earlier for the chaotic reflections near the base of the escarpment may also
explain this seismic facies. Because line GT2-24 obliquely crosses a canyon
(Figure 1), the scattering of acoustic energy along the canyons irregular
topography also appears to be the most likely interpretation.

DISCUSSION
The erosional truncation of inferred Early Cretaceous toe-of-slope facies
by the MCU, and apparent truncation of platform interior facies along the
escarpment, both indicate that the southern Florida escarpment was eroded
from the early or middle Cenomanian through the late Paleocene. The
maximum amount of erosion can be measured by using the distances
between the most basinward point of erosional truncation by the MCU and
the present base of the escarpment, and the angle at which the seismic lines
cross the inferred Early Cretaceous platform margin trend (Figures 3-6). On
line GT2-11, 6.6 km of erosion has occurred, whereas on line GT2-24, 6.1
km of erosion is evident. These distances agree with the estimate of 5 to 10
km made by Freeman-Lynde (1983), which he based on inferred widths of
depositional facies belts (e.g., Wilson, 1975; Read, 1985).
Assuming that the most basinward point of truncation by the MCU

represents the maximum amount of erosion at the base of the escarpment,


and that bypass slopes developed during Early Cretaceous time, we have
reconstructed the platforms depositional slope just prior to the MCU (early
or middle Cenomanian; Figures 3-6). When our reconstructions are
projected onto depth sections of seismic lines across the southern Florida
escarpment, we will then be able to estimate the volume of Lower
Cretaceous rocks that were eroded from this escarpment. Preliminary
calculations indicate that approximately 800 km3 of Lower Cretaceous
rocks were eroded from the escarpment between lines GT2-11 and GT2-24.
The canyon development associated with the eroded southern Florida
escarpment points to a cause-and-affect relationship, suggesting that
localized, point-source mechanisms (e.g., turbidity currents or debris flows)
might have been more important than more laterally effective processes
(e.g., contour-following currents). Alternatively, Faust (1984) identified
extensive truncation but very few basinward-trending channels on the MCU
surface seaward of the Florida escarpment, implying that contour-following
currents might also have affected the base of the escarpment.
At nearby DSDP site 540 (Figure 1), the Late Cretaceous hiatus that
corresponds to the MCU is represented by 56 m of sediment gravity flows,
which contain a normal stratigraphic succession of early or middle
Cenomanian, Maastrichtian, and late Paleocene biota (Buffler et al., 1984).
Schlager et al. (1984) proposed that platform drowning, current scouring,
and slumping on tectonically steepened slopes generated this thin Late
Cretaceous section. We think that similar mechanisms resulted in the
erosion of the southern Florida escarpment. Based on all the available data,
we envisage the following scenario for the erosion of the southern Florida
escarpment.
During the early or middle Cenomanian, there was a rapid drop and
subsequent rise in relative sea level, probably the cumulative result of
tectonic movements (i. e. tilting the platform eastward) and eustatic sealevel fluctuations. Subaerial erosion along the top of the platform margin
combined with submarine gravity flows began eroding the Early Cretaceous
platform slope. Large-Scale slumping and a reorientation of the paleoslope
angle, both associated with tectonic movements, might also have
contributed to the erosional process. Contour-following currents probably
reworked debris accumulating at the base of the slope, further contributing
to its erosion. Since the MCU corresponds with a Late Cretaceous hiatus, it
is likely that erosion continued through the Late Cretaceous. The intensity
of gravity flows probably diminished through time because the Early
Cretaceous margin was drowned and the shelf-edge stepped-back toward
the Florida Peninsula.
There does not appear to be any post-MCU erosion along the base of the
escarpment on lines GT2-11 and GT2-24 because there is no apparent
truncation of the MCU by younger unconformities (Figures 3-6). The MCU
can be traced to just below the present base of the escarpment and is

Erosion of the southern Florida escarpment

150

overlain by thin wedges of younger strata (Figures 3-6). Large-scale


Tertiary and Quaternary slumps are observed on the slope above the
escarpment (Mullins et al., 1986; Twichell et al., 1986). It is possible that
the declivity of the escarpment is so great that material eroded from above
the escarpment bypasses it and accumulates at its base without further
eroding the thick Lower Cretaceous section exposed along the escarpment.
Presently, post-MCU, Texas Marathon Oil Company, Littleton, Colorado
NOPEC Geo Services, A.S., Oslo, Norway Petrobras, Rio De Janeiro,
Brazil Petro-Canada Resources, Calgary, Alberta, Canada Sarawak Shell
Berhad, Sarawak, Malaysia Schlumberger-Doll Research, Ridgefield,
Connecticut Scripps Institution of Oceanography, La Jolla, California Shell
Oil Company, Houston, Texas Texas A & M University, College Station,
Texas Total-CFP (Compagnie Francais des Petroles), Paris, France
Universite Bordeaux I, Institut de Geol headward erosion might be
occurring in the many canyons that incise the southern Florida escarpment,
although their initial formation was probably associated with the MCU
erosional period.

ACKNOWLEDGMENT
Support for this research was provided by funding from a National
Science Foundation grant (OCE-8417771) to R.T. Buffler. We thank R.
Freeman-Lynde and C. Paull for discussions and reviews of early drafts of
this manuscript.

REFERENCES CITED
Addy, S.K., and R.T. Buffler, 1984, Seismic stratigraphy of the shelf and slope,
northeastern Gulf of Mexico: AAPG Bulletin, v. 68, p. 1782-1789.
Antoine, J., R. Martin, T. Pyle, and W. Bryant, 1974, Continental margins of the
Gulf of Mexico: in C. Burk and C. Drake, eds., Geology of Continental
Margins: New York, Springer-Verlag, p. 683-694.
Bryant, W., A.A. Meyerhoff, N. Brown, M. Furrer, T. Pyle, and J. Antoine, 1969,
Escarpments, reef trends and diapiric structures, eastern Gulf of Mexico:
AAPG Bulletin, v. 53, p. 2506-2542.
Buffler, R.T., J. Watkins, J. Worzel, and F.J. Shaub, 1980, Structure and early
geologic history of the deep central Gulf of Mexico: in R. Pilger; ed.,
Proceedings of Symposium on the Origin of the Gulf of Mexico: Baton
Rouge, Louisiana State Univ., p. 3-16.
Buffler, R.T., et al., 1984, Initial Reports of the Deep Sea Drilling Project, v. 77:
Washington, D.C., U.S. Govt. Printing Office, 747 p.
Corso, W., and R.T. Buffler, 1985, Seismic stratigraphy of Lower Cretaceous
carbonate platforms and margins, eastern Gulf of Mexico (abs.): AAPG
Bulletin, v. 69, p. 246.
Faust, M.J., 1984, Seismic stratigraphy of Middle Cretaceous Unconformity in the
central Gulf of Mexico basin: Unpubl. M.A. Thesis, Univ. of Texas at
Austin, 164 p.

W. Corso, R.T. Buffler and J.A. Austin, Jr.

Freeman-Lynde, R.P., 1983, Cretaceous and Tertiary samples dredged from the
Florida escarpment, eastern Gulf of Mexico: Gulf Coast Association of
Geological Societies Transactions, v. 33, p. 91-99.
Garrison, L., and R. Martin, 1973, Geologic structures in the Gulf of Mexico basin:
United States Geological Survey, Professional Paper 773, 85 p.
Locker, S., and R.T. Buffler, 1983, Comparison of Lower Cretaceous carbonate
shelf margins northern Campeche Escarpment and northern Florida
Escarpment: in A. Bally, ed., Seismic Expression of Structural Styles: v. 2,
p. 2.2.3-1232.2.3-128.
Mitchum, R., 1978, Seismic stratigraphic investigation of West Florida slope, Gulf
of Mexico: in A. Bouma, G. Moore, J. Coleman, eds., Framework, Facies
and Oil-Trapping Characteristics of the Upper Continental Margin: AAPG
Studies in Geology 7, p. 193-224.
Mullins, H.T., A.F. Gardulski, and A.C. Hine, 1986, Catastrophic collapse of the
west Florida carbonate platform margin: Geology, v. 14, p. 167-170.
Paull, C.K., B. Hecker, R. Commeau, R.P. Freeman-Lynde, A.C. Neumann, W.
Corso, S. Golubic, J. Hook, E. Sikes, and J. Curray, 1984, Biological
communities at the Florida escarpment resemble hydrothermal vent taxa:
Science, v. 226, p. 965-967.
Phair, R., 1984, Seismic stratigraphy of the Lower Cretaceous rocks in the western
Straits of Florida, southeastern Gulf of Mexico: Unpubl. M.A. Thesis,
Univ. of Texas at Austin, 319 p.
Read, J.F., 1985, Carbonate platform facies models: AAPG Bulletin, v. 69, p. 1-21.
Schlager, W., and R. Ginsburg, 1981, Bahama carbonate platformsthe deep and
the past: Marine Geology, v. 44, p. 1-24.
Schlager W., R.T. Buffler, D. Angstadt, R. Phair; 1984, Geologic history of the
southeastern Gulf of Mexico: in R.T. Buffler et al., Initial Reports of the
Deep Sea Drilling Project, v. 77: Washington, D.C., U.S. Govt. Printing
Office, p. 715-738.
Shaub, F.J., R.T. Buffler, and J. Parsons, 1984, Seismic stratigraphic framework of
deep central Gulf of Mexico basin: AAPG Bulletin, v. 68, p. 1790-1802.
Twichell, D.C., L.M. Parson, P.C. Valentine, and C.E. Paull, 1986, Long-range
side-scan sonar survey of eastern Gulf of Mexico (abs.): AAPG Bulletin, v.
70, p. 657.
Vail, P.R., R.M. Mitchum, Jr., and S. Thompson, III, 1977, Seismic stratigraphy
and global changes of sea level, Part 4: Global cycles of relative changes in
sea level, in C.E. Payton, ed., Seismic StratigraphyApplications to
Hydrocarbon Exploration: AAPG Memoir 26, p. 83-97.
Wilson, J.L., 1975, Carbonate Facies in Geologic History: New York, SpringerVerlag, 471 p.
Worzel, J., et al., 1973, Initial reports of the Deep Sea Drilling Project, v. 10:
Washington, D.C., U.S. Govt. Printing Office, 748 p.

APPENDIX
ACQUISITION AND PROCESSING INFORMATION
LINE: AG-4, TIME SECTION
Energy Source: 1500-cubic-inch airguns (4) fired at 500 PSI
Number of channels: 48
Group Interval: 300 feet, 91.4 meters
Minimum Offset Distance: 875 feet, 267 meters
Maximum Offset Distance: 7775 feet, 2370 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 24 Fold)
5) Bandpass Filter: 5/10-30/60 Hz
6) Mute to Water Bottom
7) AGC (500ms)
LINE: SF-9, TIME SECTION
Energy Source: 1500-cubic-inch airguns
(4) fired at 500 PSI
Number of channels: 48
Group Interval: 230 feet, 70 meters
Minimum Offset Distance: 871 feet, 265.5 meters
Maximum Offset Distance: 11,666 feet, 3555.5 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 24 Fold)
5) Bandpass Filter: 5/10-40/50 Hz
6) Mute to Water Bottom
7) AGC (500ms)

LINE: G2-11, TIME SECTION


Energy Source: 1500-cubic-inch airguns
(4) fired at 400 PSI.
Number of Channels: 24
Group Interval: 300 feet, 91.4 meters
Minimum Offset Distance: 688 feet, 209.7 meters
Maximum Offset Distance: 7588 feet, 2312.7 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 12 Fold)
5) Time variable Bandpass Filters
Water Bottom: 5/10-55/70 Hz
1.0 s below Water Bottom: 4/8-32/48 Hz
7.0 s below Water Bottom: 3/6-24/36 Hz
6) Mute to Water Bottom
7) AGC (500ms)
LINE: GT2-1, TIME MIGRATED SECTION
Energy Source: 1500-cubic-inch airguns
(4) fired at 400 PSI
Number of Channels: 24
Group Interval: 300 feet, 91.4 meters
Minimum Offset Distance: 688 feet, 209.7 meters
Maximum Offset Distance: 7588 feet, 2312.7 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 12 Fold)
5) Predictive Deconvolution
6) Time Migration
7) 3-Trace Weighted Mix
8) Time variable Bandpass Filters
Water Bottom: 5/10-55/70 Hz
1.0 s below Water Bottom: 4/8-32/48 Hz
7.0 s below Water Bottom: 3/6-24/36 Hz
9) Mute to Water Bottom
10) AGC (500ms)

Erosion of the southern Florida escarpment

151

LINE: GT2-24, TIME SECTION


Energy Source: 1500-cubic-inch airguns
(4) fired at 400 PSI
Number of Channels: 24
Group Interval: 300 feet, 91.4 meters
Minimum Offset Distance: 710 feet, 216.4 meters
Maximum Offset Distance: 7610 feet, 2319.4 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 12 Fold)
5) Time Variable Bandpass Filters
Water Bottom: 5/10-55/70 Hz
1.0 s below Water Bottom: 4/8-32/48 Hz
7.0 s below Water Bottom: 3/6-24/36 Hz
6) Mute to Water Bottom
7) AGC (500ms)
LINE: GT2-24, TIME MIGRATED SECTION
Energy Source: 1500-cubic-inch airguns
(4) fired at 400 PSI
Number of Channels: 24
Group Interval: 300 feet, 91.4 meters
Minimum Offset Distance: 710 feet, 216.4 meters
Maximum Offset Distance: 7610 feet, 2319.4 meters
1) Demultiplex
2) Sort
3) Velocity Analyses (Semblance)
4) Normal Moveout/Stack (Nominal 12 Fold)
5) Predictive Deconvolution
6) Time Migration
7) 3-Trace Weighted Mix
8) Time Variable Bandpass Filters
Water Bottom: 5/10-55/70 Hz
1.0 s below Water Bottom: 4/8-32/48 Hz
7.0 s below Water Bottom: 3/6-24/36 Hz
9) Mute to Water Bottom
10) AGC (500ms)

W. Corso, R.T. Buffler and J.A. Austin, Jr.

Erosion of the southern Florida escarpment

152

SEISMIC STRATIGRAPHY OF CARBONATE


PLATFORM SEDIMENTS, SOUTHWEST FLORIDA

D. BRADFORD MACURDA, JR.


The Energists, Houston, Texas
The seismic-stratigraphic exploration of carbonate sediments must be
conducted within a different conceptual framework than that of siliciclastic
sediments. First, the carbonate sediments are produced in the basin in which
they are deposited, rather than transported from an external source. Second,
the higher velocity of carbonate sediments, 5000 to 7000 m/sec (16,000 to
22,000 ft/sec) decreases the resolution recorded by an individual reflector.
Third, the susceptibility of carbonate sediments to secondary processes
(such as recrystallization, dolomitization, karstification, and solution), and
the importance of these in creating reservoirs, reorients the focus of the
seismic-stratigraphic investigation.
The West Florida shelf and slope have been an area of carbonate
deposition from Early Cretaceous time onward (Mitchum, 1978). Seismic
lines of a conventional horizontal scale shot in the platform environment
(Figure 1) are difficult to interpret in the deeper parts of the record section
because reflectors either have the appearance of the proverbial straight-line
railroad tracks, or lack continuity altogether (e.g. from 1.5 to 2.0 sec;
Figure 1).
If there are any multiples present, they tend to parallel the primary
reflectors and it is difficult to differentiate between them. The use of a
squeezed section (Figure 2) overcomes the problem of recognizing the
external configuration of the sequences and where the sequence boundaries
are. Figure 2 (as originally plotted) is a 1:6 compression of the original full
line from which Figure 1 was extracted. Figure 1 comes from the upper left
end of the line seen in Figure 2 (note break in slope). The latter line was
created by summing each of the three adjacent traces and then dropping
alternate groups to produce a 1:6 compression.
In Figure 2 it is obvious that there are interpretable data down to below
2 sec (travel time) in the line. The effects of the karstification of the
Miocene carbonates also are very evident, with prominent collapse features
apparent on the left of the section from 0.5 to 1.0 sec.
In Figure 1, several sequences can be seen that show downlap to the
southwest, recording offbank transport of the carbonate sediments. Droxler

D.B. Macurda, Jr.

and Schlager (1985) investigated this phenomenon in the Bahamas and


documented the export of carbonate sediments to deeper water wherever the
platform tops are flooded and produce sediment. As they noted, this is
opposed to the response of siliciclastic margins, where sediment is stored on
the inner shelf during highstands and then transported into deeper-water
environments (such as the continental rise and abyssal plain) during
lowstands. Thus, the seismic stratigraphic investigation of carbonate shelves
and the timing of offbank transport provide a sensitive barometer for
understanding the position of sea level relative to the carbonate margin.

REFERENCES CITED
Droxler, A.W., and W. Schlager, 1985, Glacial versus interglacial sedimentation
rates and turbidite frequency in the Bahamas: Geology, v. 13, no. 11, p.
799-802.
Mitchum, R.M., Jr., 1978, Seismic stratigraphic investigation of West Florida
slope, Gulf of Mexico: AAPG Studies in Geology 7, p. 193-223.

Carbonate platform sediments, southwest Florida

159

SEISMIC STRATIGRAPHY OF THE EXMOUTH PLATEAU

R.D. ERSKINE
Exxon Production Research Company
and
P.R. VAIL
Rice University, Houston, Texas
INTRODUCTION
The Exmouth Plateau is located oceanward of Australias northwest
shelf, in water depths ranging from 800 to 2000 m. It is a sunken crustal
block about 150,000 km2, flanked by the Argo, Gascoyne, and Cuvier
abyssal plains to the north, northwest, and southwest, respectively.
Bounding margins consist of steep escarpments that slope toward oceanic
crust. The plateau is separated from Australia to the southeast by the
Exmouth subbasin/Kangaroo syncline.
Stratigraphically, the section on the platform consists of a very thick,
nonmarine to marginally marine pre-Rhaetic Triassic section. It is overlain
by a thin, marine Triassic Rhaetic and Jurassic section representing a time
of slow deposition. The wells drilled by the Esso/BHP Group suggest that
the Triassic is at least 2000 m thick on the plateau, and Barber (1982)
suggests that it may be more than 4000 m thick based on the results of wells
drilled by Phillips Petroleum north of the Esso acreage. The thin Jurassic
section in turn is overlain by a Berriasian-Valanginian-age clastic wedge
that progrades from southeast to northwest on the plateau and exceeds 1500
m in thickness. A thin, Hauterivian-Aptian-age section, consisting of
glauconitic sands on the shelf within the study area, overlies the thicker
wedge. The overlying Cretaceous Aptian-Tertiary section consists of finegrained, deep-marine marls.

DATA BASE
The data base consists of a 1978-vintage grid of approximately 10,000
km of excellent quality seismic data shot and processed jointly by Esso
Exploration and Production Australia, and BHP Petroleum Pty., Ltd., along
with logs from eight wells drilled by this group in 1979 and 1980. Previous
studies in the literature provide background information on Exmouth

R.D. Erskine and P.R. Vail

Plateau. The physiography was described by Falvey and Veevers (1974).


The geology of Exmouth Plateau was discussed by Willcox and Exon
(1976), Exon and Willcox (1980), and Von Stackelberg et al. (1980), and
many others.

INTERPRETATION AND APPLICATION OF CONCEPTS


The Norian through mid-Eocene section was interpreted in a seismic
stratigraphy study, indicated by the stippled area shown in Figure 1. Special
emphasis was placed on the Early Cretaceous wedge. The age model shown
is based upon tying the paleontologic data from the wells to regional
unconformities identified seismically, and integrating this information with
the Global Cycle Chart shown in Figure 2 (from Vail et al., 1977, 1984;
Todd and Mitchum, 1977; Mitchum, 1985; and Mitchum and Uliana, 1985).
The Berriasian-Valanginian wedge seen on lines 71 and 109 (Figures 3 and
4) was of particular interest and generally can be described seismically as
having a complex sigmoid-oblique geometry (Mitchum et al., 1977) with
the 132 Ma. unconformity as its base. Line 72 (Figure 5) demonstrates
erosion associated with this unconformity. The excellent data quality in the
area allowed this wedge to be subdivided into a number of individual,
smaller-scale units identified by reflection terminations. On line 71 a series
of basinally restricted, seismically mounded (both internally and externally)
units can be seen. These units are interleaved with, and pinch out at the toes
of, the prograding wedges. Mitchum (1985) first presented this line and
interpretation. The surfaces defining the tops of the individual wedges are
numbered 1 through 8 and the associated basinally restricted units are
numbered 1A through 8A. These restricted units thin and onlap out to the
southwest and are not seismically present on line 109.
Four wells penetrate the interval of interest within the study area:
Investigator 1, Sirius 1, Vinck 1, and Eendracht 1. Both the Eendracht 1 and
Vinck 1 wells penetrated basinal shales and claystones within the wedge
interval and will not be considered further here.
The Sirius and Investigator wells are shown on the cross section in
Figure 6 and are located on seismic lines 71 and 109 (Figures 3 and 4). The
labeled horizons on the cross section correspond to the same horizons
carried on the seismic lines with the data being integrated via synthetic
seismograms. Notice how applying the geometries of the seismically

correlated horizons to the logs allows recognition of the relationships of the


sands in the two wells. The sandstones above the 132 Ma. unconformity in
the Investigator well were tied to the basinally restricted units which pinch
out at the base of the preceding slope, and are interpreted as deep-water
deposits. The sands in the upper part of the wedge in the Sirius well are
interpreted as being deposited in shallow water at or near base level, and are
time-equivalent to the shales deposited on the slope of the clinoforms
penetrated by the Investigator well.

APPLICATION OF SEQUENCE STRATIGRAPHY TO


DEPOSITIONAL MODEL
The depositional model shown in Figure 7 schematically illustrates the
relation of sea level and subsidence to stratal patterns and facies. This model
is explained in detail by Posamentier and Vail (in preparation). Horizon 1 is
interpreted as a Type 1 unconformity due to the erosion seen on the shelf
and slope, and the occurrence of lowstand fans above it in the basin. This
surface is interpreted as being 132 Ma. in age.
Horizon 1A is interpreted as the top of the lowstand fan systems tract
and consists of massive, blocky sands in the Investigator well. The canyon
cut into the shelf seen on line 72 (Figure 5) may have acted as a conduit for,
and possibly a source of, the lowstand fan sediments.
This systems tract can be seismically subdivided into an upper and
lower unit as shown on line 71 (Figure 3). The lower unit consists of a
seismic facies that is internally mounded (nearly chaotic) with some bidirectional downlap and bow tie features, and becomes one-cycle-thick
basinward. Mutti (1985) describes three types of turbidite depositional
systems that might be related to the lowstand-fans seen here. His Type 1
(canyon detached) system consists of large, widely distributed flows that are
unchanneled. The lower unit seen here is widely distributed and the nearly
chaotic nature of the mounded area might indicate that this facies does not
represent channel complexes in this particular unit, and therefore might be
similar to Muttis Type I.
The upper lowstand-fan unit displays the same seismic facies as the
lower unit, with the mounded facies being less chaotic. This unit also
pinches out seismically, a little further updip than the preceding unit.
Muttis Type II (canyon attached) system consists of a channelized

complex grading laterally into thick-bedded sandstone lobes and


subsequently a thin-bedded lobe fringe. This system appears to be similar to
the upper unit seen here. No Mutti Type III leveed channel system was
interpreted here. In other areas the Mutti Type III turbidite is interpreted to
be the lower part of the lowstand-wedge systems tract.
Units 2 through 8 are interpreted as the lowstand-wedge systems tract,
with units 2A through 8A being basin-floor turbidites possibly derived from
slumping of the prograding shelf sands. Some of these turbidites exhibit a
seismically mounded facies, whereas others are predominantly one cycle
thick. These turbidites consist of thinner-bedded sands and shales in the
Investigator well. If they are in fact derived from slumping of the
prograding wedges rather than from a process related to a major fall of sea
level and canyon-derived, then they are different from Muttis (1985) three
types of turbidites and could represent a fourth type. Though not
schematically represented in Figure 7, these turbidites are interleaved with
the toes of the units making up the prograding complex of the lowstandwedge systems tract. Normally the presence of these lowstand-wedge
turbidites would not be predicted. However, here the high sedimentation
rate and high percentage of sand causes these turbidites to develop and be
distributed over a wide area. The quality of the seismic data and the very
low interval velocities of the sands allow seismic resolution. The absence of
sand on the slopes of the clinoforms in either well, and the seismic pinchout of the lowstand-fans and lowstand-wedge turbidites at the toe of the
slope, supports the interpretation that the lowstand-fan and lowstand-wedge
turbidites are discrete units from the prograding wedges and have no timeequivalent facies on the shelf. The toplap and associated shallow-water
sands within the prograding complex of the lowstand-wedge systems tract
indicate that maximum regression occurred at the top of this systems tract.
The downlap surface represents a time during which we interpret
deposition to have shifted landward during the maximum rise of sea level,
resulting in slow rates of sedimentation or nondeposition. Note that this
horizon ties to a thin, high-velocity, calcite-cemented sandstone in the Sirius
well that is expressed seismically as a high-amplitude, black reflector. We
interpret this to be submarine cementation that occurred during the time of
nondeposition. This surface also is very irregular and wavy on the shelf, as
seen in Figures 3 and 4. This may be the expression of large-scale sand

Seismic stratigraphy of the Exmouth plateau

163

waves having formed in relatively shallow water depths, although there is


no direct evidence for this. No transgressive systems tract is interpreted
within the study area.
The next highstand systems tract subsequently prograded basinward
until 129 Ma., when sea level again fell and another Type 1 unconformity
was formed based upon the presence of a lowstand-fan encountered in the
Scarborough well above this sequence boundary. A Type 2 sequence
boundary occurs at 127 Ma. and the overlying shelf margin wedge is
displayed on both line 71 and 109. This prograding complex was penetrated
by the Investigator well at its shelf margin and consists of shales at the base,
coarsening upward into a sandstone at the top. No lowstand-fan is
associated with the Type 2 sequence boundary.

ACKNOWLEDGMENTS
The results presented here primarily are from work done at Exxon
Production Research Company (EPR) from 1982-1983 by the authors, W.B.
Harris (University of North Carolina at Wilmington), M.L. Hoerster
(formerly of EPR), E. Klein (formerly of Esso Exploration Production
Norway), and T. Bergland (Norsk Hydro). P.R. Vail collaborated with Esso
Australia in the pre-drill exploration phase. J. Hardenbol and L.E. Stover of
EPR aided in the construction of the age model. Special thanks to Esso
Exploration and Production Australia and BHP Petroleum Pty., Ltd., for
permission to publish the data.

Mutti, E., 1985, Turbidite systems and their relations to depositional sequences, in
G.G. Zuffa, ed., Provenance of arenites: NATO Advanced Study Institutes
Series, Series CMathematical and Physical Sciences, 148, p. 65-93.
Todd, R.G., and R.M. Mitchum, Jr., 1977, Seismic stratigraphy and global changes
of sea level, part 8: identification of Upper Triassic, Jurassic, and Lower
Cretaceous seismic sequences in Gulf of Mexico and offshore West Africa,
in C.E. Payton, ed., Seismic stratigraphyapplication to hydrocarbon
exploration: AAPG Memoir 26, p. 145-164.
Vail, P.R., 1987, Seismic stratigraphy interpretation procedure, in A.W. Bally, ed.,
Atlas of seismic stratigraphy: AAPG Studies in Geology 27, v. 1, p. 1-10.
Vail, P.R., R.M. Mitchum, Jr., and S. Thompson III, 1977, Seismic stratigraphy and
global changes of sea level, part 4: global cycles of relative changes of sea
level, in C.E. Payton, ed., Seismic stratigraphyapplication to
hydrocarbon exploration: AAPG Memoir 26, p. 83-97.
Vail, P.R., J. Hardenbol, and R.G. Todd, 1984, Jurassic unconformities,
chronostratigraphy, and sea level changes from seismic stratigraphy and
biostratigraphy, in J.S. Schlee, ed., Interregional unconformities and
hydrocarbon accumulation: AAPG Memoir 36, p. 129-144.
Von Stackelberg, U., N.F. Exon, U. Von Rad, P. Quilty, S. Shafik, H. Beirsdorf, E.
Seibertz, and J.J. Veevers, 1980, Geology of the Exmouth and Wallaby
Plateaus off northwest Australia: sampling of seismic sequences: Bureau
Mineral Resources Journal of Australian Geology and Geophysics, v. 5, p.
113-140.
Willcox, J.B., and N.F. Exon, 1976, The regional geology of the Exmouth Plateau:
Australian Petroleum Exploration Association Journal, v. 16, p. 1-11.

REFERENCES CITED
Barber, P.M., 1982, Paleotectonic evolution and hydrocarbon genesis of the central
Exmouth Plateau: Australian Petroleum Exploration Association, v. 22, pt.
1, p. 131-144.
Exon, N.F., and J.B. Willcox, 1980, The Exmouth Plateau: stratigraphy, structure,
and petroleum potential: Australian Bureau of Mineral Resources Bulletin
199, 58 p.
Falvey, D.A., and J.J. Veevers, 1974, Physiography of the Exmouth and Scott
Plateaus, western Australia and adjacent northeast Wharton basin: Marine
Geology, v. 17, p. 21-59.
Mitchum, R.M., Jr., P.R. Vail, and J.B. Sangree, 1977, Seismic stratigraphy and
global changes of sea level, part 6: stratigraphic interpretation of seismic
reflection patterns in depositional sequences, in C.E. Payton, ed., Seismic
stratigraphyapplications to hydrocarbon exploration: AAPG Memoir 26,
p. 117-134.
Mitchum, R.M., Jr., 1985, Seismic stratigraphic expression of submarine fans, in
O.R. Berg and D. Woolverton, eds., Seismic stratigraphy IIan integrated
approach: AAPG Memoir 39, p. 117-137.
Mitchum, R.M., Jr., and M.A. Uliana, 1985, Seismic stratigraphy of carbonate
depositional sequences, Upper Jurassic-Lower Cretaceous, Neuquen basin,
Argentina, in O.R. Berg and D. Woolverton, eds., Seismic stratigraphy II
an integrated approach: AAPG Memoir 39, p. 255-274.

R.D. Erskine and P.R. Vail

Seismic stratigraphy of the Exmouth plateau

164

CENOZOIC CARBONATE BANKS, FOZ DO


AMAZONAS BASIN, NORTHEASTERN BRAZIL

M.J.N.BROUWER
and
M.M. SCHWANDER
Koninklijke/Shell Exploratie en Produktie Laboratorium
INTRODUCTION
The Foz do Amazonas basin forms part of the Northeastern Brazilian
passive margin and is associated with the Neogene Amazon delta. This
basin is flanked to the northwest by the Precambrian Guiana shield and to
the southwest by the Precambrian Brazilian shield. It is separated by the
Gurupa arch from the intracratonic Paleozoic Amazon basin (Figures 1 and 2).
The Foz do Amazonas basin contains in excess of 6000 m of Cretaceous
and Cenozoic sediments. Its evolution is closely associated with the
transform opening of the Equatorial Atlantic Ocean. The Neogene clastic
fan of the Amazon delta prograded onto oceanic crust. The continentaloceanic boundary is located (approximately) under the present-day shelf
break.

GEOLOGICAL EVOLUTION
During the Early Cretaceous rifting phase, continental clastics
accumulated in wrench-induced bains and pull-apart grabens, as evident in
the nearshore parts of Figure 2. In the Foz do Amazonas basin, marine
conditions were established during the Albian and Cenomanian early phases
of sea-floor spreading in the Equatorial Atlantic (Figure 3). The area of the
Foz do Amazonas basin remained tectonically active during the earlier parts
of the Late Cretaceous. This tectonic activity can be related to transform
motions along the St. Pauls and Romanche fracture zones, which
apparently projected under the continental margin of northeastern Brazil.
With progressive opening of the Equatorial Atlantic, the area of the Foz
do Amazonas basin became tectonically quiescent and its further evolution
was governed by lithospheric cooling, contraction, and sediment loading.
Late Cretaceous and Cenozoic sediments formed a seaward, expanding
wedge.
From the Paleocene to early Miocene, the clastic supply to this passive

M.J.N. Brouwer and M.M. Schwander

margin basin was limited. On the outer shelf this is reflected in the
development of a mixed carbonate-clastic platform. Landward, these
carbonate platforms are offset by generally fine, low-energy terrigenous
clastics. During the late Miocene, the Amazon drainage system developed
as a result of the synorogenic uplift of the Andes. This induced a massive
influx of clastics onto the Foz do Amazonas shelf, causing a sharp
termination of the carbonate-dominated depositional regime.
During the late Miocene to Holocene, the shelf areas subsided only a
little and consequently were bypassed by the bulk of the clastic load of the
Amazon River. This induced the outbuilding of the submarine Amazon cone
over oceanic crust. Large-scale slope instability beyond the earlier
carbonate shelf edge was associated with growth faulting and clay
diapirism.

PALEOCENE TO EARLY MIOCENE CARBONATE


PLATFORMS
Basically, four carbonate cycles are recognized, corresponding to the
Paleocene, late Eocene, Oligocene, and early Miocene. Examples of the
seismic expression of late Eocene to early Miocene carbonate banks are
given in Figures 5 and 6.
Carbonate banks generally are associated with moderately continuous,
moderate-to-low-amplitude reflections. The tops of individual carbonate
banks, corresponding to depositional sequence boundaries, are expressed by
continuous, relatively high-amplitude reflections (see well 1-APS-33E,
Eocene-Oligocene boundary). The interbank deposits consist of shales,
marls, and minor carbonate stringers and give rise to laterally continuous
high-amplitude reflections.
Three types of carbonate bank margins are recognized: 1. Relatively
abrupt carbonate bank margins, generally facing the open sea, are
characterized by short, discontinuous, dipping reflection patterns which are
interpreted as peri-platform carbonate talus. This seismic facies is, however,
not calibrated by the available well control.
2. A gradational transition between the carbonate bank facies and the
interbank facies characterizes the landward side of the Oligocene platforms,
calibrated by well 1-APS-33E.
3. The early Miocene carbonate platform appears to prograde to the

north over Oligocene interbank deposits. Early Miocene interbank deposits


are typified by an onlap-offlap reflection geometry and probably contain
carbonate debris (storm deposits?).
The boundary between middle and late Miocene sediments corresponds
to a regionally correlative marker that is associated with the contact of
carbonates and overlying prodelta clays.
The late Miocene depositional sequence consisting of shales and minor
sands accumulated under shallow marine conditions and reflects the vertical
aggradation of prodelta and deltaic complexes. This series is characterized
by parallel reflection geometries; its lower, low-reflective part is clay-prone
whereas its upper, highly reflective part contains frequent sand stringers.
The upper boundary of this sequence is erosional. Overlying PliocenePleistocene deposits generally are characterized by chaotic to paralleldiscontinuous reflection patterns, and probably consist of several cut-andfill depositional sequences that cannot be resolved in detail on the basis of
the available data.

Petroleum Geology, v. 4, p. 3-34.


Kumar, N., 1981, Geologic history of the North and Northeastern Brazilian
margin; controls imposed by seafloor spreading on the continental
structures: Canadian Society Petroleum Geologists Memoir 7, p. 527-542.
Mabesoone, J.M., J.J. Fulfaro, and K. Suguio, 1981, Phanerozoic sedimentary
sequences of the South American platform: Earth Science Review, v. 17, p.
49-67.
Nairn, A.E.M., and F.G. Stehli, 1973, A model for the South Atlantic, in A.E. M.
Nairn and F.G. Stehli, eds., The ocean basins and margins, v. 1, the South
Atlantic: New York, Plenum Press, p. 1-24.

ACKNOWLEDGMENT
We gratefully acknowledge the permission for publication granted by
Enserch International Exploration, Inc., Dallas; Pecten International
Company, Houston; Petrobras Supex, Rio de Janeiro; Marathon Petroleum
Amazonas, Ltd., Houston; and Shell Internationale Petroleum Mij., V.B.,
The Hague.

REFERENCES CITED
Asmus, H.E., and F.C. Ponte, 1973, The Brazilian marginal basins, in A.E.M.
Nairn and F.G. Stehli, eds., The ocean basins and margins, v. 1, the South
Atlantic: New York Plenum Press, p. 87-137.
Bigarella, J.J., 1973, Geology of the Amazon and Parnaiba basins, in A.E.M. Nairn
and F.G. Stehli, eds., The ocean basins and margins, v. 1, the South
Atlantic: New York, Plenum Press, p. 25-86.
Campos, C.W.M., F.C. Ponte, and K. Miura, 1974, Geology of the Brazilian
continental margin, in C.A. Burk and C.L. Drake, eds., The geology of
continental margins: New York, Springer-Verlag.
Carozzi, A.V., 1981, Porosity models and oil exploration of Amapa carbonates,
Paleogene, Foz do Amazonas basin, offshore NW Brazil: Journal

Cenozoic carbonate banks, Foz Do Amazonas basin

174

PARA-MARANHAO BASIN, BRAZIL

PETROBRAS EXPLORATION DEPARTMENT


The Para-Maranhao basin is located on the equatorial Brazilian margin,
and covers an area of approximately 90,000 km2. The stratigraphy of the
basin can be summarized as consisting of a basal clastic section of
Cretaceous age, overlain by a Tertiary carbonate platform, which can reach
up to 6000 m in thickness.
Seismic stratigraphic interpretation defines three major sequences
related to phases of basin evolution. These sequences are divided into
several smaller seismic units, and they are shown in profiles A and B.
Sequence I, of Aptian age, comprises sediments deposited during the rift
phase of the basin. It is divided into two units: Unit 1 corresponds to
lacustrine and fluvial sediments; and Unit 2 corresponds to sediments
deposited during the first marine transgressions in the basin.
Sequence II comprises sediments deposited during the beginning of the
thermal subsidence phase (postrift) of the basin. It also is divided into two
units. Unit 3 is of Albian age, and corresponds to sediments deposited in
restricted and shallow-marine environments. Seismically, this unit presents
an onlap pattern indicative of transgressive sea conditions. Unit 4 is of
Turonian/Santonian age, and corresponds to a thick clastic sequence
deposited on a slope environment. It is defined on the seismic profiles by a
conspicuous onlap pattern.
Sequence III comprises sediments deposited in the thermal subsidence
phase of the basin. It is divided into six units. Unit 5 is
Campanian/Maastrichtian. This unit corresponds to clastic sediments
deposited on a slope environment. It also contains carbonates which mark
the beginning of formation of a major carbonate platform on the shelf. Unit
6 and Unit 7 date from the Paleocene to early-to-middle Eocene. These two
units have a similar nature, and correspond to a period of intense carbonate
deposition. Prograding ramps originating an offlap pattern characterize units
6 and 7. Unit 8 is middle-to-late Eocene; Unit 9 is Oligocene; and Unit 10 is
Miocene to Present. These three units are similar in nature, and correspond
to shelf carbonates seismically defined by horizontal reflectors.
The area was strongly affected by structures associated with listric faults
that cut across the entire section (profile B).

REFERENCES
Asmus, A.E., 1976, Conhecimento atual da margem continental Brasileira: Rio de
Janeiro, Petrobras Internal Report.
Ojeda, H.A.O., 1982, Structural framework, stratigraphy and evolution of Brazilian
marginal basins: AAPG Bulletin, v. 77, p. 732-749.

Figure 1. Location map of Para-Maranhao basin, offshore Brazil, including an index showing
profiles A (Figure 2) and B (Figure 3).

Petrobras Exploration Department

Para-Maranhao basinBrazil

179

POTIGUAR BASIN, BRAZIL

PETROBRAS EXPLORATION DEPARTMENT


The Potiguar basin is located on the equatorial northeastern coast of
Brazil, and its offshore portion covers an area of approximately 16,000
km2. Sediments deposited during the Early Cretaceous consist of alluvial
fans and fluvial deltaic lacustrine sequences. These were followed, during
the Albian-Aptian, by siliciclastic and transitional deposits, which were
partially eroded. Above the unconformity are carbonate platforms of Late
Cretaceous age, and prograding clastic marine sediments, which correspond
to Upper Cretaceous and Tertiary deposits. Part of the basin has been
affected by volcanismrepresented by basaltic flows during the Early
Cretaceous and the Tertiary.
Two seismic profiles, located in the central part of the basin illustrate
some stratigraphic features.

REFERENCES
Araripe, P.T., 1985a, Nota tecnica, DIRNOE/SECEPO: Rio de Janeiro, Petrobras
Internal Report.
Araripe, P.T., 1985b, Revisao da coluna estratigrafica da Bacia Potiguar, in Semana
de Estudo sobre a Fanerozoico Nordestino, Recife.
Asmus, A.E., 1976, Conhecimento actual da margem continental Brasileiro: Rio de
Janeiro, Petrobras Internal Report.
Ojeda, H.A.O., 1982, Structural framework, stratigraphy, and evolution of
Brazilian marginal basins: AAPG Bulletin, v. 77, p. 732-749.
Souza, S.M., 1982, Atualizacao da litoestratigrafia da Bacia Potiguar: 21st
Congress Brazilian Geology (Salvador), v. 5, p. 2392-2406.

SEISMIC PROFILE A
Three major seismic stratigraphic sequences are identified in this
profile, and they represent a deltaic sequence (II), bounded by two
carbonate platforms (I and III).
Sequence I was deposited during the Albian-Cenomanian, and has an
average thickness of 200 m in this area. Internal reflections show
predominantly parallel configuration.
Sequence III is Turonian-Santonian in age, and has an average
thickness of 400 m. Internal reflections also are mostly parallel.
Normal faults cut across the entire section, and the terminations of the
carbonate sequencesas seen in this profilesupport two different
interpretations. One is erosion after deposition of Sequence III, and the
other is retrogradation of both platforms.

SEISMIC PROFILE B
This profile is similar to profile A, and is approximately parallel to it.
Sequence III was apparently affected by an adiastrophic fault, which caused
sliding of a block along the slope. Internal reflections in the slided block dip
opposite to the regional dip, indicating rotation of the block or local section
growth.

Petrobras Exploration Department

Figure 1. Location map of Potiguar basin, offshore Brazil, showing locations of profiles A
and B (Figures 2 and 3).

Potiguar basinBrazil

185

JEQUITINHONHA BASIN, BRAZIL

PETROBRAS EXPLORATION DEPARTMENT


The Jequitinhonha basin is one of the sedimentary basins on the eastern
Brazilian coast. Its evolution comprises three phases: interior fracture,
interior sag, and marginal sag.
The interior fracture phase evolved during the Neocomian, and
sediments were deposited in fluvial-deltaic-lacustrine sequences.
The next phase, interior sag, evolved during the Aptian. At this time,
subsidence created an elongated depression that was invaded by the sea;
representative sediments consist mostly of evaporites (anhydrite and halite)
and clastics.
Progressively, a narrow sea was created where fan deltas were
interbedded with limestones on a broad shelf of Albian-Cenomanian age.
With increasing oceanic opening and seaward tilting of the basin the shelf
was covered by Late Cretaceous slope and basinal shales (marginal sag
phase).
Finally, in the Tertiary, an offlapping sequence prograded over the
onlapping slope. During the early Tertiary (Paleocene to upper Eocene)
shales were the predominant sediments, whereas in the middle Tertiary
(Oligocene to Miocene) carbonate sequences dominated sedimentation on
both shelf and slope.
The Jequitinhonha basin lacks Jurassic sediments, which are present in
other Brazilian marginal basins. The basin was affected by volcanism,
which is represented by wide basalt flows, during much of its history.
The seismic profile shown is located in the center of the basin, with a
northeast-to-southwest orientation. Ages and lithologies are indicated. Of
particular interest are the intense faulting of the Albian carbonate platform,
and the sharp transition from clastic to carbonate facies in the Tertiary.

REFERENCES
Asmus, A.E., 1976, Conhecimento actual da margem continental Brasileira: Rio de
Janeiro, Petrobras Internal Report.
Dauzacker, M.V., H. Schaller, A.C.M. Castro, Jr., and M. Marroquim, 1984,
Geology of Brazils Atlantic margin basins: Oil and Gas Journal (March
4), p. 142-144.
Nascimento, M.M., K. Tsubone, and M.B. Araujo, 1984, Arcabouco tectonoestrutural simplificado e classificacao das areas prospectaveis das Bacias
do Espirito Santo, Mucuri, Cumuruxatriba e Jequitinhonha: Rio de Janeiro,
Petrobras Internal Report.
Ojeda, H.A.O., 1982, Structural framework, stratigraphy and evolution of Brazilian
marginal basins: AAPG Bulletin, v. 77, p. 732-749.

Figure 1.Location map showing the area of the Jequitinhonha basin, offshore Brazil, as well as the
orientation of the seismic profile in Figure 2.

Petrobras Exploration Department

Jequitinhonha basinoffshore Brazil

191

OUTCROP MODELS FOR SEISMIC STRATIGRAPHY:


EXAMPLES FROM THE TRIASSIC OF THE DOLOMITES

A. BOSELLINI
Istituto di Geologia
University of Ferrara, Ferrara, Italy
GEOLOGIC SETTING
This contribution shows examples of large-scale stratigraphic
relationships exposed in the Triassic of the Dolomites (Northern Italy;
Figure 1). Discussion focuses on progradation geometries of carbonate
platforms and the platforms relationships with adjacent sediments; all the
features shown are at a scale (500 to 1000 m in thickness) comparable with
those found in seismic profiles.
There are two groups of prograding platforms in the Triassic of the
Dolomites (Figure 2). The older group is early Ladinian (about 230 to 235
m.y.); the younger one is early Carnian (about 225 to 230 m.y.). Their
basinal counterparts are, respectively, the Livinallongo and San Cassiano
formations.
The two carbonate periods are separated by an important magmatic
and tectonic event, which affected the central area of the Dolomites.
Associated with the intense mafic volcanism and intrusions are welldocumented extensional and compressional structures of Middle Triassic
age.
In the late Carnian, a different carbonate system, the Durrenstein
Formation, developed in the central-eastern Dolomites. This new system,
which grew by an aggradational mechanism, infilled the remnant, shallow
Cassian basins and, accreting vertically, onlapped the flanks of the previous
platforms.
The deposition of a thin, shaly, terrigenous, and varicolored
successionthe Raibl Formationmarks the end of a period of most varied
topography and sedimentation.

THE LADINIAN MODEL


The Ladinian platforms represent the deeply indented edge of a wider
carbonate shelf located to the west and to the south of the Dolomites proper.
Some small isolated platforms or buildups (Sassolungo, Sella, Latemar,

A. Bosellini

Cernera, etc.), however, grew independently in front of this edge. Platforms


and buildups prograded almost horizontally over the adjacent, starved
basins (Livinallongo Formation; Figures 3-A, 3-B, 3-C), but the amount of
progradation (1-6 km) was quite different from place to place, even in the
same buildup. Progradation was simultaneous with aggradation, as shown
by the thick (600 to 800 m), horizontally bedded sediments of the platform
interior. The upper boundary is characterized by offlap relationships
between the horizontally bedded platform interior and the clinoform flank
deposits (Figure 3-D). Quite often the facies change is very abrupt.
Progradation of the Ladinian platforms was stopped by the late Ladinian
magmatic and tectonic events. Where buried (fossilized) by volcanic
deposits, the original platform morphology is preserved.
The Ladinian model (Figure 3) implies a relative rise of sea level
(tectonic subsidence in our case) during platform development and
progradation. The Ladinian platforms and the coeval basin sediments
represent a depositional sequence bounded by unconformities and their
correlative conformities. The downlap surface (or condensed sequence), on
which the platform prograded, is the basinal Livinallongo Formation. The
resulting carbonate body is tabular.

THE CARNIAN MODEL


During the entire middle to late Ladinian, platforms were secondary
features relative to the major geological events of tectonic collapse, volcanic
activity, and deformation. These platforms did not continue to subside and,
in many cases, were even subaerially exposed with consequent vadose
diagenesis and karstification. In Carnian time, once the subaerial volcanic
remnants were eroded so that they no longer supplied detritus, carbonate
sedimentation resumed and new carbonate platforms developed. These new
buildups and platformsunlike their Ladinian precursorsdeveloped on an
irregular and rough submarine morphology (maximum depth of the sea
bottom was approximately 500 to 600 m).
Normally, the Carnian platforms nucleated (grafted) onto older
Ladinian ones and expanded laterally, prograding over the basinal San
Cassiano Formation (a thick volcaniclastic and turbiditic succession), and
became shallower upward into neritic conditions.
Accordingly, the following Carnian model of carbonate platform

evolution and progradation can be delineated (Figure 4). As shown by


toplap relationships and by absence of internal horizontally bedded facies,
Carnian progradation (Figures 4-A, 4-B, 4-C, 4-D, 4-E) took place during a
relative stillstand of sea level. All Carnian platforms nucleated and started
their progradation necessarily from elevated areas (i.e., the older, exposed,
and karstic, Ladinian platforms). Having to face a high basin-sedimentation
rate, they commonly show a climbing lower boundary (Figure 4-A). The
Carnian platforms, therefore, thin outward and develop a typical wedge
shape.

CESSATION OF PROGRADATION
The cessation of progradation is documented where the platforms are
fossilized by younger rocks onlapping their flanks. Among the variety of
phenomena that stopped the advance of the Triassic platforms, two cases of
well-exposed relationships are fossilization by volcanics and fossilization
by carbonates (Figure 5). The original platform-basin morphology is
preserved in several places by a thick cover of volcanic rocks. The volcanics
wedge out against the slopes of the buildups (Figure 5-A), covering them
completely in places. The late San Cassian morphology was buried by a
subtidal carbonate succession, the Durrenstein Formation, which onlaps the
flanks of the buildups by wedging against their slopes (Figure 5-B). These
peculiar stratigraphic relationships are better explained if, assuming a
eustatic lowering of sea level, the Carnian platforms were subaerially
exposed, killed, and eroded; meanwhile, some San Cassian basins were,
by then, so shallow that a lowering of sea level by 50 m (or so) could have
triggered carbonate sedimentation on their floors. The Durrenstein gradually
accreted, first filling the last of the San Cassian basins, then onlapping the
flanks of the carbonate buildups, and, finally, partly covering them.

Bosellini, A., 1984, Progradation geometries of carbonate platforms: examples


from the Triassic of the Dolomites, northern Italy: Sedimentology, v. 31, p.
1-24.
Bosellini, A., and D. Rossi, 1974, Triassic carbonate buildups of the Dolomites,
northern Italy, in L.F. Laporte, ed., Reefs in time and space: SEPM Special
Publication 18, p. 209-233.
Castellarin, A., and G.B. Vai, eds., 1982, Guida alla Geologia del Sudalpino
centro-orientale: 1 centenario della Societa Geologica Italiana, Guide
geol. reg.: Bologna, Societa Geologica Italiana, 381 p.
Cros, P., 1974, Evolution sedimentologique et paleostructurale de quelques plateformes carbonatees biogenes (Trias des Dolomites italiennes): Sciences
Terre, v. 19, p. 299-379.
Doglioni, C., 1986, Tectonics of the Dolomites (Southern Alps, Northern Italy):
Journal Structural Geology, in press.
Fursich, F.T., and J. Wendt, 1977, Biostratinomy and palaeoecology of the Cassian
Formation (Triassic) of the Southern Alps: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 22, p. 257-323.
Gaetani, M., et al., 1981, Nature and evolution of Middle Triassic carbonate
buildups in the Dolomites (Italy): Marine Geology, v. 44, p. 25-57.
Leonardi, P., et al., 1967, Le Dolomiti. Geologia dei monti tra Isarco e Piavetwo
volumes: Rome, Cons. Naz. Ricerche, 1019 p.
Wendt, J., 1982, The Cassian Patch Reefs (Lower Carnian, Southern Alps): Facies,
v. 6, p. 185-202.

REFERENCES
Biddle, K.T., 1979, Characteristics of Triassic carbonate buildups of the Dolomite
Alps, Italy: evidence from the margin-to-basin depositional system: Rice
University, PhD thesis, Houston, 216 p.
Blendinger, W., 1986, Isolated stationary carbonate platforms: the Middle Triassic
(Ladinian) of the Marmolada area, Dolomites, Italy: Sedimentology, v. 33,
p. 159-183.

Outcrop models for seismic stratigraphy

194

REGIONAL SEISMIC STRATIGRAPHIC ANALYSIS OF UPPER JURASSICLOWERCRETACEOUS CARBONATE


DEPOSITIONAL SEQUENCES, NEUQUEN BASIN, ARGENTINA

ROBERT M. MITCHUM, JR.


and
MIGUEL A. ULIANA*
Exxon Production Research Company
Houston, Texas
*Present address: Astra Cia. Argentina de Petroleo S.A.,
Buenos Aires, Argentina
INTRODUCTION
We intend to demonstrate how vintage seismic and well data were
combined into a regional stratigraphic study in an area where modern-day
shooting generally was not available to us. The regional seismic data
included mostly single-fold variable-density film shot in 1962, although
some modern multi-fold data were made available. The lesson to be learned
is that old data can be very useful in regional studies, especially when
combined with well information and some modern data.
The units studied are the Upper Jurassic-Lower Cretaceous Loma
Montosa, Quintuco, and Vaca Muerta formations, which are important
reservoirs in the Neuquen basin, Argentina. Our investigations used the
seismic stratigraphic technique outlined in Vail et al. (1977, part 7) to
interpret an open regional grid of reflection profiles provided by Esso
Exploradora y Productora Argentina and Yacimientos Petroliferos Fiscales.
Because data from 65 wells and numerous outcrops were available, we
incorporated additional steps to use this information.
A more complete account of the study and a comprehensive
bibliography may be found in Mitchum and Uliana (1985).

REGIONAL SETTING
Figure 1 shows the location of the Neuquen basin in west-central
Argentina, some of the producing fields, and the position of seismic lines,
wells, and outcrops referred to in this chapter. The Neuquen is a
predominantly Jurassic-Cretaceous rifted depositional basin. It is roughly

R.M. Mitchum, Jr. and M.A. Uliana

Figure 1. Neuquen basin location map.

triangular, with the southern and northeastern depositional edges meeting in


an acute angle opening to the west. Sediments thicken to the west and
northwest, but the basin is terminated along its western boundary by northsouth Tertiary structures of the Andes Mountains.
From a tectonic standpoint, the basin lies in a continental intraplate
setting, just east of the belt of pronounced Andean deformation.
Sedimentation in the area was triggered by the inception of a Late Triassic
rift system characterized by large, tilted blocks. The rift fill is Early Jurassic
to Kimmeridgian and is made of marine and non-marine clastics, with
subordinate carbonates and evaporites (Figure 2).
During the Late Jurassic and Early Cretaceous, the basin shifted into an
early sag phase, and a thick series of carbonates, shales, and sandstones

Figure 2. Regional stratigraphy of Neuquen basin.

accumulated. Tithonian, Berriasian, Valanginian, and Hauterivian rocks,


including those discussed in the present chapter, are referred to as the Vaca
Muerta, Quintuco, Mulichinco, and Agrio formations. Subsidence histories
for Neuquen wells suggest a slow and constant rate of thermal subsidence
for this time, which is the interval of interest in this chapter. Siliciclastic
sediment influx also was fairly constant and moderately low, as indicated by
abundant autochthonous carbonate deposits, including oolites, formed under
narrowly defined environmental conditions.
In the late Cretaceous, a thick unit of red beds was deposited in a broad,
poorly defined basin during a late sag phase. The remainder of the
Figure 3. Tithonian-Valanginian lithostratigraphic units.

Carbonate depositional sequences, Neuquen basin

206

stratigraphic column, not illustrated in Figure 2, consists of discontinuous


Tertiary rocks coeval with the Andean deformation to the west.
The Neuquen basin is the second most prolific hydrocarbon producer in
Argentina (Figure 1). Almost all the potential reservoirs produce
hydrocarbons somewhere in the basin, but the Jurassic Lotena-Punta Rosada
and Tordillo-Sierras Blancas sandstones (Figure 2), and the Upper JurassicLower Cretaceous Quintuco-Loma Montosa carbonates are the major
reservoirs (Figure 3). Los Molles and Vaca Muerta black shales are
widespread source rocks. Jurassic block faulting, wrenching, drape folding,
and superposed Andean deformation provide structural and stratigraphic traps.

SEISMIC STRATIGRAPHIC INTERPRETATION

Figure 4. Regional seismic cross section of Tithonian, Berriasian, and Valanginian sequences plotted in depth
from seismic section along line A-A (see Figure 1). Generalized lithologies from wells are tied to section.

Figure 4 is a regional geologic cross section made from a key seismic


section. Generalized lithologies and depth conversions are based on wells
shown on the section and in Figure 5. The seismic section is shown in
Figures 6 and 7. The section (A-A in Figure 1)is about 250km (155mi) in
length. Tithonian, Berriasian, and Valanginian rocks have been subdivided
into a series of nine depositional sequences (units A-J) making up the Vaca
Muerta, Quintuco, Loma Montosa, and Mulichinco formations. Sequence
boundaries are assigned dates (m.y. before present) based on the time scale
used by Haq, Hardenbol, and Vail (1987). The section was converted to
depth using wells tied to the seismic data, and a datum is set on the top of
the Valanginian. In general, the top and base of the overall TithonianValanginian interval are slightly divergent because of thickening from
southeast to northwest. The lateral changes in lithology within each of the
sequences are schematically indicated in Figure 4 and will be discussed
further in the following sections.
Each sequence is broadly lenticular, with a thick central portion flanked
by much thinner but continuous wedges that gradually became thinner in
updip and downdip directions. The internal-reflection configuration of these
sequences defines a typical clinoform depositional pattern. As shown in
Figure 4, the thick portions of the successive sequences tend to be displaced
laterally in a strongly prograding pattern of basin fill. The thick portions of
the sequences have the most reflection terminations and are the best places
to define the sequence boundaries. Away from these critical positions, the
boundaries were correlated through the seismic grid, mostly by tracing the
physical continuity of essentially parallel reflections.

CORRELATION OF DEPOSITIONAL SEQUENCES


IN WELLS
Figure 5. Regional geologic cross section of Tithonian, Berriasian, and Valanginian stages, using the
top of Valanginian as datum. Seismic sequence boundaries are correlated in wells. Wells are
located along B-B (see Figure 1).

R.M. Mitchum, Jr. and M.A. Uliana

Sequence boundaries defined on seismic data also were identified and


correlated in nearby wells, using a combination of velocity, biostratigraphy,
and lithology. Figure 5 is an east-west well section, located along line B-B
(Figure 1), which closely parallels the seismic cross section of Figure 4. The

Carbonate depositional sequences, Neuquen basin

207

Figure 8. Correlation of Neuquen basin sequences with coastal onlap and eustatic
sea-level chart (Haq, Hardenbol, and Vail, 1987); and correspondence of geometry of
sequences with trends of eustatic changes.
Figure 10. Depositional model of Tithonian-Valanginian depositional sequence.

nine wells on the cross section are tied to the seismic section, and seismic
sequences are identified in the wells. The section datum is on the uppermost
seismic surface (top of unit J), which in most cases is the top of the
Mulichinco Formation and is a good marker in the wells. This uppermost
unit is a series of continuous sands of relatively constant thickness.
The lower, major boundary is the base of the Vaca Muerta Formation,
which is a good marker point in all the wells, marking the base of the
characteristic basinal dark-gray organic shale of this formation. The upper
boundary of sequence F is the best-correlated intermediate surface. It is
characterized by a persistent limestone bed, which, although it changes in
texture across the basin, provides a good marker through the middle of the
unit. Its clinoform characteristics are evident in the westernmost four wells,
where the marker drops about 250 m (820 ft) in the section from east to
west.
Figure 4 illustrates the time-transgressive nature of the lithostratigraphic
units. Because given lithologies tend to occupy similar environmental
settings in each successively younger sequence, the lithostratigraphic units
tend to become younger from east to west. For example, the top of the Vaca
Muerta Formation, which consists mostly of dark organic basinal shales, is
early Tithonian in age in eastern areas; westward it becomes as young as

R.M. Mitchum, Jr. and M.A. Uliana

Figure 9. Seismic section along line C-C. See Figure 1 for location.

early Valanginian. The diachronous nature of the formations can be seen in


the regional cross sections of Figure 4.

REFLECTION CONFIGURATION
Figures 6 and 7 show the western and eastern parts, respectively, of the
seismic section used in Figure 4 (section A-A, Figure 1). This section, shot
in 1962, is single-fold, variable-density film, which was state-of-the-art at
that time. Although data quality leaves much to be desired, the basic

reflection configurations can be interpreted adequately for regional work.


Sequence F (Figure 6) shows a good example of reflection terminations
(primarily onlap and truncation, secondarily downlap, and toplap) that
define the boundaries of the sequence. The thick part of this unit has a
prograding complex-oblique internal configuration, which changes rapidly
into mostly parallel updip and downdip components. Thin updip segments
of the overlying sequences (G, H), and downdip segments of the underlying
sequences (A-D, E), occur above and below sequence F. A minor unit (G)

appears to be a restricted lowstand wedge (Posamentier and Vail, 1987)


onlapping the downdip edge of the thick portion of F. The clinoform
reflections within and at the upper surface of sequence F show a prominent
change of slope that is interpreted as a shelf edge. The segment updip of this
change is interpreted as shelf environment, and the downdip segment as the
slope and basin environments of sequence F. Lithologic and biostratigraphic
data from wells tied to the seismic section support this interpretation.
Similar interpretations are made for other sequences in Figures 6 and 7.

Carbonate depositional sequences, Neuquen basin

210

sequence F (Figure 6). Values for older sequences such as D or C are less
than 200 m (656 ft) and even smaller for sequences A and B (Figure 7).

DATING SEISMIC SEQUENCES AND CORRELATING


THEM TO GLOBAL EVENTS

Figure 12. Thickness map (contoured in two-way seismic time) and depositional
environments of sequence B.
Figure 13. Thickness map (contoured in two-way seismic time) and depositional
environments of sequence F.

Figure 11. Shelf margins of Tithonian-Valanginian depositional sequences.

Updip portions of sequences (shelf) are thin (30 to 80 m; or 98 to 262


ft), widespread, and characterized by mostly parallel, concordant,
continuous high- to medium-amplitude reflections. The overall thinning of
these members is very gradual and occurs partly through successive onlap
terminations of the lowermost reflections as well as by convergence of the
individual beds.
Central portions of sequences (slope) occupy the generally thicker (100
to 200 m; or 328 to 656 ft) prograding lenses between the thin updip and
downdip segments, as typified in sequence F.
Downdip portions of sequences (basinal) generally consist of thin (10 to
60 m, or 33 to 197 ft) distal clinoform toes of parallel to convergent, fairly
continuous, high- to medium-amplitude reflections with low-angle
downlapping to concordant patterns. Although the downdipping clinoform
toes may appear to terminate seismically by downlap, electric log

R.M. Mitchum, Jr. and M.A. Uliana

correlation shows the basinal toes to be fairly continuous for long distances,
with thicknesses below seismic resolution. Similarly, ammonite zonation in
outcrops tends to confirm the existence of thin black shales spread over the
entire distal realm of the basin.
The Tithonian-Valanginian sequences show changes in the overall
geometry of the prograding units and can be subdivided into three groups of
patterns (Figure 8). Sequences A to C show dominantly sigmoidal
configurations, poorly defined shelf-slope breaks, and fairly thick,
widespread, updip (shelf) segments. In these units, toplap configurations are
minor and restricted to the youngest parts of the sequences. Sequences D to
F have much more pronounced shelf margins, complex-oblique internal
configurations, and very thin, updip (shelf) components. Toplap
configurations are prominent and tended to form throughout the
development of the unit. Sequences G and H have pronounced shelf-slope

breaks and very thick but areally restricted updip shelf segments. We think
the differences in configuration of these three groups of sequences are due
mostly to differences in rates of sea-level rise, but other actors could be
related, such as greater water depths controlled by faster subsidence in
western basin areas and more carbonates in sequences A to C compared
with more clastics and texturally different carbonates in younger sequences.

PALEOWATER DEPTH ESTIMATES


An estimate of the minimum paleowater depth in the basin near the
slope break at the time of each sequence can be made by measuring the
relief between the slope break and the base of the slope portion of the
clinoform, along a single clinoform reflection which approximates a
continuous depositional surface. Using this procedure, we estimated a
minimum basinal paleowater depth of 250 m (820 ft) at the time of

Estimated ages of the nine sequences identified in the present study are
shown in Figure 8. Sequences A to C are thought to be Tithonian; sequence
D straddles the Tithonian-Berriasian boundary; E and F are Berriasian; G
and H are considered Valanginian; and I and J (Mulichinco Formation) are
thought to be depositional units in the lower part of the latest ValanginianHauterivian sequence.
This estimate is based on comparison of Neuquen basin sequences with
the coastal onlap and sea-level cycle chart of Vail et al. (1977) and Haq et
al. (1987). Although no faunal data were available in study wells, a general
Tithonian-Valanginian age for this interval was established by comparison
to the outcrop, where detailed ammonite zonation studies were done
(Leanza, 1973; Leanza and Hugo, 1977).
The lower boundary of the overall Vaca Muerta-through-Mulichinco
interval (base of sequence A) is a very distinctive seismic sequence
boundary and was established as the base of Tithonian both on outcrops and
in wells. The global coastal-onlap chart of Figure 8 shows that the 138-m.y.old surface (near base of Tithonian) is a prominent cycle boundary on the
chart and probably correlates with the lower boundary of the interval.
The most prominent and widespread internal sequence boundary within
the interval is the top of sequence F. This sequence extends relatively farther
onto the shelf than the restricted sequences above it. This relationship is
best seen in the seismic section of Figure 9, which is a north-south section
located on the southern flank of the basin (Figure 1, section C-C ). The
upper boundary of sequence F is a prominent, well-defined, continuous
reflection, and the three sequences above it onlap against it to the south
(left), indicating a restricted distribution. Sequence F and older sequences
maintain fairly constant thicknesses southward, implying a relatively
widespread shelfward extent.
On the global coastal-onlap chart (Figure 8), the most pronounced
seaward shift in coastal onlap within the interval of interest is located at the
128.5 Ma. cycle boundary (latest Berriasian). This surface coincides in time
with the boundary between the widespread sequence F and the restricted
sequence G-G.
Sequences below sequence F are assigned Tithonian and Berriasian
ages. There is a close coincidence between the local sequences and the
global chart for these ages, as shown in Figure 8. Sequences above sequence
F are interpreted as latest Berriasian to Valanginian in age. Unit G is
interpreted as the lowstand wedge (Posamentier and Vail, 1987) of sequence
G. Both G and H pinch out against sequence F (Figure 9). Unit I is

Carbonate depositional sequences, Neuquen basin

211

restricted to the shelf margin and basinward and may represent the lowstand
wedge of a sequence containing I and J (Mulichinco Formation). Unit J is
widespread and is interpreted as a transgressive deposit overlain by a
marked flooding surface that forms the prominent seismic reflection at the
top of the overall mapped unit, dated at 120.5 Ma. on the sea-level chart.

BASIN FILL GEOMETRY AND EUSTASY


Figure 8 shows subdivision of the sequences into three groups based on
their geometry. Sequences A-C show a sigmoid prograding-aggrading
pattern; sequences D-F have a complex-oblique, mostly prograding pattern
with prominent shelf edges; and G and H have complex-oblique patterns
with prominent aggradation and shelf margins. We interpret these
geometries to relate in part to trends of eustatic change. Sequences A-C
were deposited during a period of overall slow sea-level fall but were
interrupted by rapid falls and rises. These rapid sea-level rises coupled with
basinal subsidence might tend to drown the outer shelf, producing some
aggradation and sigmoid configuration. Sequences D-F were deposited
during a period of highstand or slow fall with only minor drops and rises.
With little accommodation on the shelf, sediments would prograde rapidly
into deeper slope areas. Sequences G and H were deposited after a major
eustatic fall during the ensuing rapid rise. This sea-level rise, coupled with
basinal subsidence, allowed accommodation on the shelf and generated
significant aggradation as well as progradation.

DEPOSITIONAL MODEL OF A SEQUENCE

SHELF MARGIN MAPS

ACKNOWLEDGMENTS

The prominent change in slope of a seismic reflection interpreted as the


shelf margin is one of the most important paleoenvironmental boundaries,
separating the shelf from the slope and basin environments (Figures 6 and
7). These boundaries can be mapped from seismic data for given sequences.
Figure 11 is a map of all the shelf margins. It gives a picture of the
progressive fill of the basin. Shelf margins of the oldest sequences (A and
B) occur farthest east and south and are closely parallel to the southern and
northeastern edges of the basin. They form a rather acute angle in the
eastern corner of the basin. Progressively younger shelf margins occur
farther northward and westward, until the youngest shelf margins have only
a slightly concave outline on Figure 11.
Thickness maps of individual sequences give patterns similar to those of
the shelf margins because the shelf margin of a given sequence occurs near
the axis of thickest deposition of the sequence. Figures 12 and 13 are
thickness maps (contoured in two-way seismic time) of sequences B and F,
respectively. The narrow zone of maximum thickness corresponds closely to
the trace of the corresponding shelf margin (Figure 11). Distribution of
interpreted shelf, slope, and basin environments are also shown.

We thank Esso Inter-America, Inc. (now Exxon Company International,


Inc.) and Yacimientos Petroliferos Fiscales for the release of this report,
including seismic and well data.

REFERENCES
Haq, B.U., J. Hardenbol, and P.R. Vail, 1987, Chronology of fluctuating sea levels
since the Triassic: Science, v. 235, p. 1156-1167.
Leanza, A.F., 1973, Estudio sobre los cambios faciales de los estratos limitrofes
Jurasico-Cretacicos entre Loncopue y Picun Leufu, Provincia de Neuquen,
Republica Argentina: Revista de la Asociacion Geologica Argentina, v. 28,
p. 97-132.
Leanza, A.F., and C.A. Hugo, 1977, Sucesion de ammonites y edad de la
Formation Vaca Muerta y sincronicas entre los paralelos 35o y 40o 1.s.,
cuenca Neuquina-Mendocina: Revista de la Asociacion Geologica
Argentina, v. 32, p. 248-264.
Mitchum, R.M., Jr., and M.A. Uliana, 1985, Seismic stratigraphy of carbonate
depositional sequences, Upper Jurassic-Lower Cretaceous, Neuquen basin,
Argentina, in O.R. Berg and D.G. Woolverton, eds., Seismic stratigraphy
IIan integrated approach to hydrocarbon exploration: AAPG Memoir 39,
p. 255-274.
Posamentier, H.W., and P.R. Vail, 1987, Eustatic controls on clastic deposition, in
Sea-level changes: SEPM Special Publication, in press.
Vail, P.R., et al., 1977, Seismic stratigraphy and global changes of sea level, in
C.E. Payton, ed., Seismic stratigraphyapplications to hydrocarbon
exploration: AAPG Memoir 26, p. 49-212.

Figure 10 is a highly generalized schematic representation of a single


depositional sequence, showing the idealized succession of lithology,
carbonate texture, thickness, and environment across the sequence. The
lower part of the diagram shows distribution of source, reservoirs and seal
within the sequence. This idealized model is based on studies of individual
depositional sequences from descriptions of wells tied to seismic sections,
literature descriptions, and field observations. The best porosity occurs in
inner-shelf environments, in oolitic and bioclastic calcarenites. Source rocks
are restricted to basin to lower-slope components. Most of the sealing
members are basin-slope deposits, shelf shales, and fine-grained peritidal
rocks.

R.M. Mitchum, Jr. and M.A. Uliana

Carbonate depositional sequences, Neuquen basin

212

TERTIARY HIATUSES IN WESTERN APPROACHES

C. RAVENNE
C. MULLER
and
L. MONTADERT
Institut Francais du Petrole
Rueil-Malmaison, France

age attributed to them, and the type and changes of the seismic facies units
making them up. As we shall show in the conclusions, seismic facies testify
to major reworking processes that underscore the difficulties of using a
single technique, such as biostratigraphy or seismic reflection profiling, to
determine the scope of the hiatuses.
We shall not dwell here on the tectonic style, which was described by
Montadert et al. (1979a, b), or on the sedimentology, which was described
in Graciansky, et al. (in press).

For additional information on the specific ties between DSDP core holes
and the seismic profiles, the reader is referred to Montadert and Poag
(1985).

INTRODUCTION

HIATUSES DETERMINED BY
BIOSTRATIGRAPHYCORRELATION WITH
SEISMIC PROFILES

Profile 601
On profile 601 (Figure 4, pages 216-227), only the upper-middle Eocene
and Plio-Pleistocene sequences can be followed from one end to the other.
The Paleocene sequence is very reduced, especially at the top of the tilted
and eroded blocks (shotpoint 1500) and sometimes disappears nearly
completely (as between shotpoints 3500 to 4600).
The upper Miocene sequence can only be distinguished between
shotpoints 200 and 5200. The Oligocene sequence can only be recognized
clearly in a few locations (OC 601-4, shotpoints 2200 to 2300, 2500 to
5500). The other sequences recognized often were difficult to correlate and
to date.
The Paleocene often is characterized by a parallel reflection seismic
facies of rather high amplitude, great continuity and, except on OC 601-1
(shotpoints 100 to 600) where the seismic facies unit is rather chaotic,
reworked deposits.
The same applies between shotpoints 1400 and 1700 (part 3) and
shotpoints 5000 and 5400 (part 8).
The lower Eocene displays a low amplitude facies, sometimes
transparent throughout the start of the profile (shotpoints 1 to 2300), and
then shows a rather chaotic facies suggesting reworked deposits, including
mass slides and mass flows.
The middle Eocene, often associated with the upper Eocene and part of
the Oligocene, nearly always consists of a seismic facies unit with
disorganized, even chaotic reflections, further emphasizing the permanence
of gravitational reworking. It appears that the reflection amplitude increases
seaward. A more detailed observation in the small basins located between
the overthrown blocks (OC 601-8, shotpoints 4800 to 5400, OC 601-9, 10,

The Deep Sea Drilling Project (DSDP) has achieved major advances in
paleooceanography. The effort was focused on the history of water
circulation in the deep oceanic basins, on detailed biostratigraphy, and on
oxygen isotopes, and the like.
Another major advance concerns the paleoenvironment of continental
margins, with the recognition by Vail et al. (1977) of unconformity-bounded
sequences on seismic reflection profiles. Two important hypotheses arose
from this work: (1) major unconformities should be recognized
worldwideat least on passive margins; and (2) eustatic changes are the
major control of post-rift margin stratigraphy.
Whereas these results are universally used, especially the so-called
Vails curve, few data are available to check these hypotheses because the
seismic data and drilling data often are not published, and because DSDP
did not focus on this problem.
We present some results of a Tertiary study based on two profiles of the
Bay of Biscay, located on the margin of Goban Spur (location in Figure
1)an area characterized by a low sedimentation rate.
The study is based on seismic stratigraphy of high-resolution
multichannel seismic profiles and the biostratigraphy of DSDP holes. We
shall concentrate on the problem of unconformities and hiatuses and discuss
their resolution.
We apply the term unconformity to the boundaries of seismic sequences. They
may, or may not, correspond to hiatuses that can be defined by biostratigraphy.
We shall describe the following: (1) hiatuses identified by the
biostratigraphic study and their regional or local correlation with seismic
profiles; and (2) the different sequences distinguished on the profiles, the

C. Ravenne, C. Muller and L. Montadert

PROFILE ANALYSIS: SEQUENCE IDENTIFICATION,


DATING, CHANGES IN SEISMIC FACIES, SCALE OF
REWORKING PROCESSES

Figures 2 and 3 show the relevant stratigraphic results of Leg 80, as


reported by Snyder et al. (1985), and Montadert and Poag (1985). The only
unconformities that occur across the entire margin are those in the middle to
upper Miocene, middle Oligocene, and upper Paleocene. Much of the Upper
Cretaceous sequence is either condensed or absent.
On both seismic profiles, only a very thin sedimentary cover occurs on
tilted blocks that are related to Early Cretaceous rifting. It is extremely
difficult to pick up unconformities that can be followed along the whole
margin. Locally a large number of seismic sequences can be defined, but
only two sequence boundaries can be tentatively traced all along these
profiles: middle-upper Eocene and upper Paleocene.
Locally datable sequence boundaries can be traced only near the holes.
However, even at the location of the hole it is difficult to positively identify
such hiatuses, because of the very low sedimentation rate. At hole 548
(Figures 2 and 3), for example, the middle-upper Miocene hiatus is well
displayed. The middle-upper Eocene horizon and upper Paleocene hiatus
also can be tentatively recognized. At site 550 (Figures 2 and 3) seismic
calibration was relatively easy and the following hiatuses were determined:

middle-upper Miocene hiatus.


hiatuses and a very condensed interval ranging from
lower Miocene to lower Eocene.
upper Paleocene hiatus.

Figure 1. Index map showing locations of profiles 601 and 603, discussed in this
chapter.

Tertiary hiatuses in Western Approaches

213

shotpoints 5600 to 6800) shows the superposition of seismic facies units


with disorganized reflections, indicating the existence of different deposits
of reworked material.
The middle Eocene only exhibits organized reflectors in segment 3,
between shotpoints 1500 and 1650. The configuration of these reflectors
suggests a deposition associated with currents.
The facies of the upper sequence generally are characterized by
continuous, parallel, high-frequency reflectors in the first half of the profile

Figure 2. Biostratigraphy of DSDP leg 80. Left half represents the early Tertiary
(Paleogene) zone in columns 548-551. The right half represents late Tertiary and
Quaternary (Neogene) in same drill sites (Jar = Jaramillo event; Old = Olduvai
event). Figure after Snyder et al., (1985).

C. Ravenne, C. Muller and L. Montadert

(shot-points 1 to 5500). However, in the basin located between shotpoints


5500 and 7000, and in the deep basin (shotpoint 7000 to the end of the
profile), the reflectors again become disorganized, except in the uppermost
part of the fill.
Major gravitational reworking is demonstrated by seismic facies units
with chaotic or even disorganized appearance starting with the Paleocene
and associated with pronounced slope break (OC 601-8). These characterize
the deposition of the lower Eocene and are responsible for the deposition of

most of the upper-middle Eocene along the entire profile. These gravity
movements still affect nearly all the subsequent sedimentation going
towards the basin.
Profile 603
Profile OC 603 (Figure 5, pages 218-239), parallel to Profile OC 601,
generally shows the same changes as profile OC 601, but a larger number of

Figure 3. Summary of sediment accumulation rates and distribution of unconformities


for sites 548 to 551. A condensed interval is defined as having a sediment accumulation
rate of less than 5m/m.y. Note that the only unconformities that occur across the entire
margin are those in the middle to upper Miocene, middle Oligocene, and upper
Paleocene, although the middle part of the Upper Cretaceous is absent or highly
condensed at all four sites (after Masson et al., 1985).

Tertiary hiatuses in Western Approaches

214

tilted blocks gives rise to several perched basins.


Only the Paleocene and middle Eocene sequences are barely
identifiable, on parts 1 and 2. Noteworthy between shotpoints 8300 and
7800 is seismic facies unit A, in the form of a wedge, consisting of several
hummocky forms for which a current-related origin is assumed. The
presence of small chaotic units in this unit A suggests the presence of
channels.
Between shotpoints 7700 and 6900 (parts 2 and 3), many
unconformities exist separating many depositional sequences that are very
difficult to correlate to the sequences located on either side of the basin; this
is because of the change in the type of deposit and the very great difficulty
of separating the unconformities at the top of the tilted blocks. Nearly all
the sequences consist of seismic facies units with disorganized or even
chaotic reflections, except the Paleocene sequences and the upper
(Quaternary?) sequences. In this environment, body B appears to be
interpreted as a deep-sea fan lobe taken in a cross section, body C as a
slope-basin deposit, and body D as a suprafan part seen in a longitudinal
section.
On part 3 (and especially between shotpoints 6900 and 6500), a
superposition of many sequences could only be distinguished by the
biostratigraphic analysis of the data from hole 548 (part 5). Only in the
upper part were a number of sequences identified on seismic profiles; on the
other hand, seismic profiles helped to distinguish all the sequences in part
OC 603-2. At the top of this truncated block, the reflections are nearly all
subparallel and continuous up to shotpoint 5500, which marks the end of the
truncated unit. Only within the middle Eocene and part of the Miocene do
we see seismic facies units with disorganized reflectors. In the basin
between the boundary of this truncated unit and the intrusion located
between shotpoints 4600 and 4500 (part 6), the reflectors generally are well
organized and hence do not indicate major mass remobilizationexcept for
the middle Eocene (general character) and for most of the sequences in the
immediate neighborhood of major reliefs. The seismic facies unit of the
principal fill, from the upper Miocene to the Pleistocene, between
shotpoints 5400 and 5100, is characterized by subparallel, continuous, highfrequency reflections and suggests a turbidite deposition.
The quality of the reflectors deteriorates increasingly, going from parts 7
to 9. In part 7, a virtual alternation exists between the seismic facies units
with disorganized reflectors and those with continuous reflectors. This is
interpreted as remobilization of sediments between turbidite units. In parts 8
and 9, nearly all the sequences display disorganized (and even chaotic)
reflectors that are separated by erosional unconformities; the screened body
E illustrates the result of a recent slump. The three parts 7, 8, and 9 exhibit a
large number of sequences that are very difficult to date. An early Eocene
age can be proposed for one of the sequences, in view of its seismic
characteristics and its position.

C. Ravenne, C. Muller and L. Montadert

The sequences identified in the oceanic plain (parts 10, 11, and 12) are
well-dated thanks to the biostratigraphic analysis of the data from hole 550.
They generally consist of seismic facies units with disorganized reflectors
representing highly reworked deposits. These sequences are separated from
each other by a continuous reflector of fairly strong amplitude,
corresponding either to condensation levels or to sedimentation hiatuses
(Figure 3). These reflectors may sometimes be intersected by multiple
faults, as frequently observed for the unconformity located at the top of the
upper-middle Paleocene sequence.
The Tertiary deposits observed along the slope and in the oceanic plain
are almost always affected by reworking and mass sliding, giving the
reflector a more-or-less disorganized appearance. The single continuous
reflectors between two deposition sequences must be associated with
condensation levels or with sedimentation hiatuses.
The units with organized reflections observed in the basins located at the
upper part of the slope suggest a greater change of the material and turbidite
deposition.

CONCLUSIONS
The large number of reworking episodes existing on the slope, within
the perched basins, and on the oceanic plain, serves to differentiate many
sequencesespecially if they are separated by calmer sedimentation
episodes (distal sedimentation of turbidite currents, for example). Multiple
erosion phases occur that often reveal great variability. Therefore, it is
difficult to date the corresponding seismic sequences and consequently the
hiatuses between them. Extensive erosion recognized on the seismic
sections may have occurred during very short time intervals that are not
resolvable by biostratigraphy. On the other hand, the cover of the truncated
blocks and of a large part of the shelf can only be subdivided into a small
number of sequences, using seismic profiles. In fact, the reflectors there
generally are subparallel, continuous, and arranged in correlatable seismic
facies units. In these zones, the sequences can only be distinguished by
biostratigraphic analysis. We now discuss the input and characteristics of
each of these systems.
In our study, seismic sequences are distinguished both by
unconformities and seismic-facies changes. We must realize the enormous
difficulties in differentiating major from minor sequences and in identifying
the important hiatuses. For instance, we may have a thick seismic sequence
due (1) to a strong variation in terrigenous influx rate caused by either
orogeny or sea-level change, and emergence of a part of the shelf, erosion of
this part, and a considerable seaward transport of sediments; or due (2) to
limited areas of deposition. However, seismic data can furnish the general
environment and the precise geometry of sedimentary bodies like canyonfill structures and base-of-slope deposits.
Biostratigraphy provides the means to date and to quantify hiatuses and

to determine which are the most important hiatuses on seismic lines. Such
hiatuses often can only be identified as a conformable reflection. The hiatus
may be a nondepositional hiatus or an apparent hiatus if caused by the
limitations of seismic resolution within conformable reflections, or else
limitations of seismic resolution, which often cannot differentiate very thin
layers.
Biostratigraphy is useful for putting erosional unconformities in their
right context (e.g., inside the middle Miocene sequence). In a normal
seismic study (i.e., a nonstratigraphic seismic study) such erosional
unconformity would be taken as the base of a sequence. This is particularly
true during low sea-level periods when terrigenous influx increases and
when a canyon-channel network crosses the slope and is active in its
construction. Biostratigraphy allows discriminations between very close
hiatuses, not differentiable on seismic data. But biostratigraphy alone often
cannot resolve the general depositional environment, which (frequently) can
only be resolved by seismic profiling.
A classic seismic survey needs to point out unconformities to separate
the main sequences. However, this may be a source of error because we
need to determine the origin of these unconformities in order to distinguish
internal erosions that are located within the sequences (channel) from the
major unconformities.
A seismic-stratigraphic study allows regional correlations, but it is very
difficult to determine the duration of hiatuses because sedimentation rates
often change.
Biostratigraphy allows us to quantify hiatuses and then to differentiate
the main sequences, but biostratigraphy must be associated with a seismic
stratigraphy study that positions the wells in their overall context. For
instance, biostratigraphy can display a large hiatus that seems to be the most
important in the area, but seismic stratigraphy may show that the location of
the well is in a local canyon or on a slumped structure.
Only the combination of both tools allows the reconstruction of the
geological story.

ACKNOWLEDGMENTS
We wish to thank the Offshore Petroleum Study Committee (Comite
dEtudes Petrolieres Marines), which allowed the shooting of these profiles.

GEOPHYSICAL PARAMETERS
Energy source:
Stacking multiplicity:
Number of channels recorded:
Interval between input channels:
Minimum offset distance:
Maximum offset distance:
Deconvolution:
Frequency filtering:
Migration:

hydraulic flexichoc
24
48
25
250 m
1175 m
yes
20-160 Hz
yes

REFERENCES CITED
Graciansky, P.C. de, et al., 1985, Initial reports of the Deep Sea Drilling Project,
leg 80, part 2: Washington, D.C., National Science Foundation, 1258 p.
Graciansky, P.C. de, and C.W. Poag, 1986, Evidence for changes in Mesozoic and
Cenozoic oceanic circulation on the southwestern continental margin of
Ireland, DSDP/IPOD leg 80: Journal of the Geological Society of London.
Hailwood, E.A., W. Bock, L. Costa, P.A. Dupeuble, C. Muller, et al., 1979,
Chronology and biostratigraphy of northeast Atlantic sediments, Deep Sea
Drilling Project leg 48, in L. Montadert, et al., Initial reports of the Deep Sea Drilling
Project, v. 48: Washington, D.C., U.S. Government Printing office, p. 305-339.
Masson, D.G., L. Montadert, and R.A. Scrutton, 1985, Regional geology of the
Goban Spur continental margin, in P.C. de Graciansky and C.W. Poag,
eds., Initial reports of the Deep Sea Drilling Project, leg 80: Washington,
D.C., National Science Foundation, 1258 p.
Montadert, L., et al., 1979a, Initial reports of the Deep Sea Drilling Project, v. 48:
Washington, D.C., U.S. Government Printing office.
Montadert, L., O. de Charpel, D.G. Roberts, P. Guennoc, and J.C. Sibuet, 1979b,
Northeast Atlantic continental marginsrifting and subsidence processes,
in M. Talwani, W.W. Hay, and W.B.F. Ryan, eds., Deep drilling results in
the Atlantic Oceancontinental margins and paleoenvironments:
American Geophysical Union, p. 154-186.
Montadert, L., and C.W. Poag, 1985, Physical properties and correlation of seismic
profiles with drilling results, in P.C. de Graciansky and C.W. Poag, eds.,
Initial reports of the Deep Sea Drilling Project, leg 80: Washington, D.C.,
National Science Foundation, 1258 p.
Muller, C., 1985, Biostratigraphic and paleoenvironmental interpretation of the
Goban Spur region based on a study of calcareous nannoplankton, in P.C.
de Graciansky and C.W. Poag, eds., Initial reports of the Deep Sea Drilling
Project, leg 80: Washington, D.C., National Science Foundation, 1258 p.
Roberts, D.G., D.G. Masson, L. Montadert, and O. de Charpal, 1981, Continental
margin from the Porcupine Seabight to the Armorican marginal basin:
Second Conference on Petroleum Geology of the Continental Shelf of
Northwest Europe (proceedings), p. 455-473.
Vail, P.R., et al., 1977, Seismic stratigraphy and global changes of sea level, in C.E.
Payton, ed., Seismic stratigraphyapplications to hydrocarbon
exploration: AAPG Memoir 26, p. 49-212.

Tertiary hiatuses in Western Approaches

215

SEISMIC STRATIGRAPHIC FEATURES OF


THE PORCUPINE BASIN, OFFSHORE IRELAND

D. BRADFORD MACURDA, JR.


The Energists
Houston, Texas
The Porcupine basin underlies a prominent re-entrant in the shelf
margin off the southwestern coast of Irelandthe Porcupine Seabight. It is
a triangle-shaped rift basin, widening out to the south. The basin was
formed during the second phase of the North Atlantic rifting during the Late
Jurassic-Early Cretaceous, and there has been periodic infilling of the basin
since that time (Roberts et al., 1981). Ziegler (1982) provided a generalized
stratigraphic column for the basin and a schematic cross section. More
recently, Graciansky et al. (1985) discussed the results of Leg 80 of the
Deep Sea Drilling Project (DSDP). The DSDP drilled four holes on the
Goban Spur, which forms the far southeastern rim of the basin.
The seismic data from Porcupine basin are of excellent quality, making
it possible to effect a very detailed seismic stratigraphic and facies analysis
down to the basement. In the following discussion, we backstrip the basin
from the top down, in order to keep better track of correlative units. Each
unit is numbered from the top down in each line; the same numbers apply in
the different sections. The numbered units consist of several sequences
grouped together; each was deposited under fairly similar environmental conditions.
Figure 1 is part of the north-south line near the eastern margin of the
basin. Unit 1 is characterized by rather parallel, continuous, mediumamplitude reflectors. There are occasional submarine channels and levees in
some parts of the basin. Several sequence boundaries within unit 1 can be
interpreted basin-wide. A few high-amplitude anomalies are present that
probably represent shallow gas drilling hazards. Unit 1 formed in bathyalabyssal water depths.
Unit 2 is most characteristically developed in the eastern one-third of
the basin. It is characterized by high-angle oblique reflectors; it shows
progradation both to the west and the south (Figures 2 and 4). It is
interpreted as a high-energy, deep-water deposit. Unit 2 is dominated by an
aggradational channel in the eastern part of the basin; it is migrating toward
the basin axis. The channel characteristically has a natural levee on its
western side.

D.B. Bradford, Jr.

Units 3 and 4 are similar; they consist primarily of parallel reflectors.


Unit 4 infilled the area around the mounds (the submarine fins) of unit 5.
Unit 4 was cut by major submarine canyons, as seen in Figure 1, where the
edges are clearly erosional. Unit 3 represents a resumption of similar
depositional conditions to those of unit 4; unit 3 infilled the canyons and
smoothed out the basin floor. Units 3 and 4 are concordant over much of the
basin and are not differentiated from one another in most of the other
figures. The unconformity between 3 and 4 correspond to the late Oligocene
unconformity (Graciansky et al., 1985) on the Goban Spur. The top of Unit
3 probably corresponds to the middle-late Miocene unconformity noted by
Graciansky et al. (1985), who characterized the interval as a time of global
climatic and paleoceanographic disturbance. Units 1 and 2 are, thus,
Pliocene to Pleistocene.
Unit 5 is one of the most distinctive units in the basin. Its external
geometry is a series of constructional mounds that are sourced from various
points on the basin margin (Figures 5 and 6). Unit 5 is a series of submarine
fans that proceed from being more distal at the base to more proximal at the
top. Individual mounds become younger to the south. The reflectors tend to
be parallel in the lower part and hummocky in the upper area.
Unit 6 is a rather innocuous-appearing interval at first glance; however,
it is one of the most significant. Its amplitude is deceiving. Internally, it
shows progradation from the basin margins, so there was centripedal fill
occurring at that time (Figure 6). It is four times thicker in the north than in
the basin center to the south; the northern end of the basin shows episodic
prograding units so there was considerable bathymetric differentiation
during its accumulation. Unit 6 occurs on top of the higher-velocity,
Maastrichtian-Danian chalk. It probably is Paleocene, whereas unit 5 is Eocene.
Unit 7 is characterized by a high-amplitude reflector at its top because
of the pronounced velocity contrast with the overlying siliciclastics and the
pronounced unconformity. It commonly is about 40 msec thick. It appears
to pinch out toward the basin margins. In the eastern third of the basin there
is a series of distinctive constructional mounds that are individually a few
kilometers across in a north-south direction (Figure 3); they are more
elongate in the east-west direction (Figure 4). The mounds form a belt 25
km (16 mi) wide, which trends northwest to southeast. The reflectors
become more parallel toward the center of the basin. Unit 7 corresponds to a

series of Upper Cretaceous-Lower Tertiary Danian chalks.


Unit 8 is, in reality, a complex series of sequences, but these are not
differentiated here. They correspond to the initial postrift fill (figures 2 and
4); Graciansky et al. (1985) dated the age of the breakup unconformity on
the Goban Spur to the south as being between the Barremian (or perhaps
Aptian) and the early Albian. This age is adopted here. The uppermost
sequences in unit 8 are a series of prograding, oblique clinoforms along the
eastern margin of the basin (Figure 4). They are herein interpreted as
alluvial fans or fan deltas.
One of the distinctive features within unit 8 is a series of high-amplitude
reflectors. These are volcanics, some of which are quite extensive
exceeding 30 km (Figure 5). They commonly are parallel to the reflectors
beneath and have onlap or downlap on top of them; they are therefore
considered to be flows. A few small ones show cross-cutting relationships
and thus are dikes.
A series of rotated fault blocks form the basin margins, and there are
some synrift basins associated with these. One of the most outstanding
features of the Porcupine basin is the volcanic ridge system (unit 9), which
developed parallel to the basin axis immediately following the rifting; it is
well displayed in Figure 2. A series of these features orient northwest to
southeast and extend several tens of kilometers. Ziegler (1982) cited
Berriasian and Barremian volcanics from the Porcupine basin.
The most distinctive feature of Figure 1 is the submarine canyons that
separate units 3 and 4. Basinward, these decrease in amplitude. Units 3 and
4 are concordant throughout most of the basin.
Figure 2 is a north-south line parallel to Figure 1 and a short distance
farther out in the basin. The line starts at 2.0 sec, and is approximately 1700
m (1 mi) in water depth. The distal end of the submarine canyons can be
seen separating units 3 and 4. The prominent volcanic feature has
considerable relief on it. Inspection of the northern side shows the onlap of
reflectors from the north interspersed with the downlap of material derived
from its summit.
Figure 3 is part of the northern continuation of the line illustrated in
Figure 2. Features of particular note include the prograding clinoforms of
unit 2, with evidence of a channel on its upper surface at the left end of the
figure. Note also the evidence of older, asymmetric channels in the older

part of the sequence to the north. The carbonate mounds of unit 7 are very
evident. Some show the local derivation of debris from their summit, such
as on the south side of the mount in left-center.
Figure 4 is an east-west line in the eastern part of the basin. The
prograding clinoforms of unit 2 are very evident. Note that there are several
groups of oblique clinoforms. Unit 7 shows the carbonate mounds in
elongate cross section; note their progradation to the west. The carbonates
onlap onto, and terminate against, the prograding clinoforms to the east at
the top sequence of unit 8. Observe how the oblique reflection character
changes to shingled as the unit is traced to the west.
Figure 5 is an east-west line in the center of the basin; it begins at 1.5
sec. The mounded character of unit 5 is evident; the unit is sourced from the
north. Units 6 and 7 show little that is distinctive because the line is near the
basin center. The flow within unit 8 is very evident; note the onlap onto its
upper surface. The crest of the volcanic ridge system (unit 9) is visible at
the base of the section.
Figure 6 is a north-south line in the western part of the basin. The
submarine fan of unit 5 is superbly developed. The sequences within unit 5
show a progressive evolution from distal to proximal fan. The upper
sequences show downlap and onlap at either edge of the fan; it was sourced
from the northwest. Note the progradation of unit 6 beneath it.

REFERENCES CITED
Graciansky, P.C. de., et al., 1985, The Goban Spur transect; geological evolution of
a sediment-starved passive continental margin: Geological Society of
America bulletin, v. 96, p. 58-76.
Hubbard, R.J., J. Pope, and D.G. Roberts, 1985, Depositional sequence mapping to
illustrate the evolution of a passive continental margin, in O.R. Berg and
D.G. Woolverton, eds., Seismic stratigraphy IIan integrated approach to
hydrocarbon exploration: AAPG Memoir 39, p. 93-116.
Roberts, D.G., D.G. Masson, L. Montadert, and O. de Charpal, 1981, Continental
margin from Porcupine Seabight to the American marginal basin, in V.
Illing and G.D. Hobson, eds., Petroleum geology of the continental shelf
of northwest Europe: London, Heyden and Sons, p. 455-473.
Ziegler, P.A., 1982, Geological atlas of western and central Europe: Shell
Internationale Petroleum Mij. B.V. (distributed by Elsevier Science
Publishers, Amsterdam), 130 p.

Seismic stratigraphic features of the Porcupine basin

240

MASS SLIDES AND TURBIDITE TYPE DEPOSITS RECOGNIZED BY OFFSHORE SEISMIC PROSPECTING:
CAP FERRET DEPRESSION AND AT THE OUTCROP: GRES DANNOT SERIES

C. RAVENNE
Institut Francais du Petrole (IFP)
Rueil-Malmaison, France
M. CREMER
CNRS-IGBA
Universite de Bordeaux I
France
P. ORSOLINI
Societe Nationale Elf Aquitaine (SNEA(P))
Paris, France
and
P. RICHE
ENSPM
Rueil-Malmaison, France
INTRODUCTION
We present three high resolution seismic profiles recorded in the
proximal part of the Cap Ferret Deep Sea Fan (Bay of Biscay). We focus
only on the stratigraphic expression of the mass slides and on the
interpretation of certain facies illustrated by the field work in the Gres dAnnot series.
The detailed and regional results of the interpretation of the Cap Ferret
profiles were published by Coumes et al. (1982), Cremer (1983), and Nely
et al. (1985). The outcrop illustrations presented are the result of work
carried out by C. Ravenne, M. Cremer and P. Riche in the Gres dAnnot
region (Figure 2). The Gres dAnnot series at the type locality and in the proximal
part of the deposits was investigated extensively by D.J. Stanley (1961, 1975, 1980).
The seismic profiles are located in the upstream part of the Cap Ferret
submarine clastic unit, or what we call the Cap Ferret canyon area, which
here corresponds to a depression that extends over an area more than 50 km
long and about 30 km wide (Figure 1).
We are concerned here only with the post-Eocene filling subsequent to
the Pyrenean orogeny. This sedimentary fill is complex and was deposited
from different sources. Part of the material is derived from the upstream part

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

of the Cap Ferret depression. Another part results from mass collapses
affecting the north flank of that depression. A large third part derives from
the canyons crossing the Landais marginal plateau.

MASS SLIDES AND TURBIDITE TYPE DEPOSITS AND


CHANGES IN THE SEISMIC FACIES RELATED TO
SLOPE COLLAPSE
Profile CF 110 (2) shows a slope collapse on the north flank of the Cap
Ferret depression and also shows the deposits of the reworked material.
From upstream to downstream we distinguish the following: zone (1), the
slip plane; zone (2), a zone where the transparent seismic facies corresponds
to crushed material (refer to Vail et al., 1977, for the methodology employed
and the analysis of these profiles); zone (3), a zone with seismic facies with
rare organized reflections that may be interpreted as the superposition of
several collapses; zone (4), a zone where the seismic facies is chaotic and
corresponds to very fractured material; and zone (5), a zone where the
seismic facies displays well-organized subparallel reflections. This last zone
corresponds to the deposition of a slope base cone where all the material
derived from the collapse area has been remobilized and fluidized, thus
allowing new stratified deposition. The complete unit, from the collapse
area to the most distal deposits discernible on seismic profiles, extends over
an area about 25 km long and 15 km wide. This whole unit lies
perpendicular to the elongated depression. Most of the deposits of the north
flank of the depression result from coalescence of reworked deposits. In the
southern part of the depression, the distal deposits of these units are
reworked by submarine currents crossing this depression. Note at (6) the
location of the present main channel.
Our description of the unit described here from the north flank of the
Cap Ferret depression has been highly simplified. Many slip planes are
actually superimposed in the main collapse area. They generate small
successions of facies zones (1) to (5) of a different scale and they
intertongue vertically. This is clearly visible in zone (4), where a few
organized reflections can already be observed, and in zone (5), where
chaotic parts occur.
This type of deposit, the change in the facies, and the process involved,
are comparable to those observed in sequence II of the profiles BAC E
(Ravenne et al., 1988).

Figure 1. Location of the profiles recorded in the Cap Ferret canyon.


Figure 2. Location of outcrop photos A1-A7, Gres dAnnot region.

Mass sliding and turbidite type deposits, Cap Ferret depression

248

Profile CF 106, parts (2) and (3), show a very long succession of
deposits resulting from slope collapses (zones where a few seismic facies
units are screen-patterned, (part (1)) on the profile CF 106-2), with the
interlocking of many seismic facies units. Clearly identifiable here is the
uniform stripping of the Aquitaine shelf (PL between shotpoints 1800 and
1920) and of the beginning of the slope, a process that we believe to be
similar to the process that caused the deposition of sequence II offshore the
Bahama scarp (profiles BAC E) (Ravenne et al., 1988).
The part surrounded by a circle (2) (CF 106-1) shows the contact
between the practically synchronous deposits which, in the southwest,
originate in the Landais marginal shelf (south of the depression), and to the
northeast correspond to the deposition of material drained by the main
channels of the depression from east to west.

Interpreted seismic profile CF 106 part (3).

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Mass sliding and turbidite type deposits, Cap Ferret depression

252

THREE SEISMIC FACIES UNITS CHARACTERISTIC OF


GRAVITY DEPOSITS AND COMPARISON WITH
SIMILAR OUTCROPS
Profile CF 109, in part (2), between shotpoints 600 and 800, shows three
combinations of facies that we compare with outcrop photographs.
Description of seismic facies units
Seismic facies unit (1) lies directly on the acoustic basement. Unit 1 is
transparent and only shows a few very low-amplitude reflections. The
regional surveys mentioned above show that unit 1 lies in the prolongation
of the very distal parts of the bottomsets related to the Landais marginal
shelf. These deposits are slope-draping deposits and consist of very fine
sediments. The drape deposits are onlapped by relatively continuous highamplitude reflections, which constitute seismic facies unit (2). This unit is
eroded and covered by seismic facies unit (3), which is characterized by
chaotic reflections. Unit 3 generally is interpreted as the result of highenergy sandy deposits capable of forming reservoirs. The field surveys
allow an easy comparison with sediments deposited in a comparable
environment. Furthermore, the dimensions of the bodies identified on
analogous outcrops are compatible with the scale of our seismic
observations.

Interpreted (above) and uninterpreted seismic profile CF 109 parts (1) and (2).

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Mass sliding and turbidite type deposits, Cap Ferret depression

257

Comparison and interpretation of


seismic facies units 1 and 2
Photo A1 shows siltstones and sandstones deposited in a deep-sea fan in
water more than 1000 m deep. They onlap on blue marls whose dip is
representative of the paleotopography. The blue marls deposited here drape
the slope in a manner very similar to seismic facies unit (1) of profile CF
109-2. The sandstone beds onlap without any erosion on the blue marls, yet
they contain grains of multicentrimetric size (see description of this
conglomerate bank in Ravenne and Beghin, 1983). Photo A2 shows this
coarse-grained deposit in detail and emphasizes its thickness.
These coarse-grained layers thicken further northward. Sandstone beds
are added at the wall and at the top of the sequence as shown by Photo A3,
which is not taken in the prolongation of Photo A1, but is representative of
this development elsewhere. Three seismic facies units (1), (2) and (3), can
be seen centered on the coarse-grained layers. The visible extension of these
units is more than 20 km long and shows very uniform thickness.
Photo A4 shows the vertical arrangement of each of these
sedimentologic units, beginning with an upward-thickening series of
sandstone beds and terminating in an upward-thinning series of sandstone
beds.
The regional and sedimentological studies of the Allos sector (Photos A3
and A4) and of the Sanguiniere range were recently carried out and
interpreted (Jean, 1985; Jean et al., 1985).
The very continuous units just described are separated by silty to shalesilt levels and are potential multilayer reservoirs. It is conceivable that
conventional seismic prospecting of these units will yield high-amplitude
reflections, that are parallel and fairly continuous, very similar to those
observed in seismic facies unit (2) of profile CF 109-2. Photo Al shows that
these units may terminate in an onlap without eroding the substratum, like
the reflections of seismic facies units (2) (profile CF 109-2). Our field
observations suggest that such seismic facies units could form concealed
stratigraphic traps, if the original material contains sandstone components.

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Mass sliding and turbidite type deposits, Cap Ferret depression

260

Photo A5 shows a unit, A, lying unconformably on another unit, B. These


two units are observed on the same mountainside. Unit B corresponds to the socalled Gres dAnnot series and their upper mesosequence (Inglis et al., 1981).
The series there is less massive and the sandstone banks are smaller than at
Avalanche Mountain (Photo A3). Unit A corresponds to the blocky shales
described by Kerckhove (1969). Here, the blocky shales consist of elements
from the upper part of the Gres dAnnot series. Unit A is entirely chaotic. We
interpret the gully erosion observed as the result of enormous slope stripping
that affected the eastern margin of the receptacle of the Annot sandstones; this
slope-stripping has probably followed the early beginnings of the Alpine
movements. This erosional process is comparable to erosions caused by the
mass slidings observed on the northern parts of profiles CF 109, CF 106, and
CF 110.
The chaotic fill of unit A normally would yield only a blind or disorganized
seismic image. Erosion is in excess of 100 m. Note the analogy with seismic
facies unit (3) of profile CF 109-2. Our analogy is important, because it points
out that a seismic facies that normally would be interpreted as a high-energy
sand deposit can also be interpreted quite differently and can result from totally
reworked material, incapable of forming a reservoir. Other experiments further
support our interpretation (Ravenne and Beghin, 1983). Beghin demonstrated
that in such a context, significant erosion can only occur by mass slides or by
hydraulic jumps over slope breaks. On the other hand, a depression fill
(depression resulting from erosion) with organized reflections could form a
potential reservoir, as Photo A6 shows. The erosion probably results from a
mass slides process, but the filling took place later.

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Mass sliding and turbidite type deposits, Cap Ferret depression

262

CHANGES IN FACIES PARALLEL TO THE INFLUX


DIRECTION
Photos A7 and A8 illustrate changes from a proximal position (Photo A7 is
approximately 45 km from the canyon outlet) to a more distal position (Photo
A8 is approximately 120 km from the canyon outlet, see the situation on the
section of the geological map in the study of profiles BAC E, Ravenne et al.,
1988). This distance of 75 km is relatively small in comparison with the
dimensions of the Cap Ferret Deep Sea Fan. It corresponds approximately to
the distance between the apex of the first canyons of the Cap Ferret depression
and the outlet of this depression in the abyssal plain. However, this distance is
enough to reduce considerably the thickness of the massive sandstone
sequences (units 1 and 2 of Photo A7, for example). A few sandstone banks
still exist in Photo A8, but they rarely are thicker than 1 m.

REMARK: ANOTHER TYPE OF FRACTURESEDIMENTATION RELATION


With profile CF 110, we showed the sedimentary result of a major slope
collapse. On profile CF 109-3, between shotpoints 1600 and 1800, the vertical
succession of facies observed in the hummocky areas should be noted. These
deposits are initiated on fault scarps. They first develop into levees (two
episodes separated by a draping stratum) with downlaps toward the southeast.
These levees are covered by draping deposits and finally filled by channel-fill
series. The interlocking of the different deposits migrates laterally toward the
southeast and may be related to a large, continuous fault scarp.

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

Mass sliding and turbidite type deposits, Cap Ferret depression

263

CONCLUSIONS
The seismic profiles presented reveal the complexity of the deep-sea fan
units derived from several source areas. This applies to the sedimentary fill
of profile CF 110, where the sequence that results from a slope-mass
collapse interferes with that resulting from the main supply sources of the
depression, which are perpendicular to it. This collapse creates a body that
differs from the main unit, but that nevertheless belongs to the same major
depositional sequence. In comparison with the Gres dAnnot series (in the
broad sense), the sequence resulting from the collapse is very similar, from
the standpoint of the lateral input in a main unit (size, facies change) to the
so-called Champsaur series or to the base of the Aiguilles dArves.
We have developed the relation between three seismic facies units and
the outcrops deposited in similar environments. In terms of reservoir
potential, the two main points are:
1) The interpretation of continuous, high-amplitude reflections lying in
onlap on the lateral discontinuity. These may result from sandstone
beds, which could form excellent reservoirs.
2) The interpretation of the chaotic seismic facies unit filling a
topographic depression caused by an erosion unconformity in a deep
environment. The comparison with the shales overlying the Gres
dAnnot series suggests that, on the contrary, these units may not
offer any reservoir potential.
Geological field surveys designed to describe bodies of size compatible
with the seismic resolution contribute to better seismic interpretations and
allow a better lithological prediction of the seismic facies units.

ACKNOWLEDGMENTS
We wish to thank the Offshore Petroleum Study Committee (Comite
dEtudes Petrolies Marines), which allowed the seismic survey; the crew of
NO Resolution; and the geophysicists of IFP who helped to collect the data
and process them.
Our special thanks go to B. Petitperrin, who participated actively in the
analysis of the seismic profiles.

C. Ravenne, M. Cremer, P. Orsolini and P. Riche

REFERENCES CITED
Coumes, F., J.R. Delteil, H. Gairaud, C. Ravenne, and M. Cremer, 1982, Cap
Ferret Deep Sea Fan (Bay of Biscay): AAPG Memoir 34, p. 583-590.
Cremer, M., 1983, Approches sedimentologique et geologique des accumulations
detritiques. Leventail profond du Cap Ferret (golfe de Gascogne) - La
serie des Gres dAnnot (Alpes de Haute Provence): These Univ. Bordeaux,
I, Ed. Technip, Paris, 344 p.
Jean, S., 1985, Les Gres dAnnot au NW du Massif de lArgentera-Mercantour:
These Univ. Scient. et Medic. de Grenoble.
Jean, S., C. Kerckhove, J. Perriaux, and C. Ravenne, 1985, Un modele paleogene
de bassin a turbidites: le Gres dAnnot du NW du Massif de lArgenteraMercantour: Geologie Alpine, t. 61, p. 115-145.
Inglis, I., A. Lepvraud, E. Mossett, A. Salim, and R. Vially, 1981, Etude
sedimentologique des Gres dAnnot: Rapport IFP-ENSPM n. 29 765.
Kerckhove, C., 1969, La zone du Flysh dans les nappes de lEmbrunais-Ubaye
(Alpes occidentales): Geologie Alpine, t. 45, p. 5-204.
Nely, G., F. Coumes, M. Cremer, P. Orsolini, B. Petitperrin, and C. Ravenne, 1985,
Leventail profond du Cap Ferret (Golfe de Gascogne, France):
Reconnaissance des divers processus sedimentaires interferents grace a
une methodologie dexoploration originale: Bull. Centr. Rech. Explor.Prod. Elf-Aquitaine 9, 2, p. 253-334.
Ravenne, C., and P. Beghin, 1983, Apport des experiences en canal a
linterpretation sedimentologique des depots des cones detritiques sousmarins: Revue de lIFP, v. 38, n. 3.
Ravenne, C., R. Vially, P. LeQuellec, and P. Valery, 1988, Deep clastic carbonate
deposits of the BahamasComparison with Mesozoic outcrops of the
Vercors and the Vocontian Trough, this volume.
Stanley, D.J., 1961, Etudes sedimentologiques des Gres dAnnot et leurs
equivalents lateraus, in Revue Inst. Franc. Petrole Ann. Combust. liquides,
v. 16, p. 1231-1254.
Stanley, D.J., 1975, Submarine canyon and slope sedimentation (Gres dAnnot) in
the French Maritime Alps: IX Internat. Congr. Sedim., Nice, 129 p.
Stanley D.J., 1980, The Saint Antonin conglomerate in the Maritime Alps: a model
for coarse sedimentation on a submarine slope: Smithsonian Cont. Mar.
Sc., Washington, n. 5, 25 p.
Vail, P.R., et al, 1977, Seismic stratigraphy and global changes of sea level, in C.E.
Payton, ed., Seismic stratigraphyapplications to hydrocarbon
exploration. AAPG Memoir 26.

Mass sliding and turbidite type deposits, Cap Ferret depression

264

SUBMARINE CANYONS, CONTOURITES,


AND VOLCANICS, MOZAMBIQUE

D. BRADFORD MACURDA, JR.


The Energists
Houston, Texas
The southeastern margin of Africa is an area of intense interest to
geologists because of its relationship to Madagascar and Antarctica prior to,
and during, the breakup of Gondwana in the Mesozoic. Martin and
Hartnady (1986) recently presented a revised reconstruction of this area
using a broad variety of data. They place the breakup of the area between
southeastern Africa and East Antarctica as being between 145 and 170 Ma.
They illustrated the position of the continental margin of southeastern
Africa (their figure 8) based on the distribution of acidic volcanic rocks that
Flores (1973) documented under the coastal plain of Mozambique. These
formed during the early separation of Africa from east Antarctica, and were either
oceanic in origin or were associated with highly fractured continental crust.
Figure 1 (line 4) is a dip line in the present-day slope-rise environment
of Mozambique, oriented northwest-southeast. It is seaward of the Jurassic
continental edge. There is a prominent unconformity (2) that separates the
synrift and postrift phases of the development of the margin. The oldest
sediments on top of the unconformity are probably Neocomian. Two waterbottom multiples (3) and (4) are present in the section. There is a series of
dipping reflectors at (1) that intersect the breakup unconformity and clearly
are not a water-bottom multiple. It is suggested that these represent synrift
volcanics, either as flows or dikes that are related to the breakup of the east
African margin. Mutter (1983) illustrated very similar features from the
Voring Plateau, northwest of Norway.
This area was drilled by the Deep Sea Drilling Project. Basalts were
recovered that are related to the earliest stages of sea-floor spreading. The
basalts were of a subaerial to shallow-water origin. A relatively large
volume of basalt flows versus relatively narrow zones of dike intrusion
caused a regular tilting of the lavas toward the spreading center (Mutter et al., 1982).
In the upper-right part of Figure 1, there is a large mound (5). On first
inspection, a slump origin might be suspected. However, the upper-righthand (southeast) side of the mound shows erosional truncation. Internally, it
shows apparent progradation to the northwest and onlap onto the rise. It
probably is a contourite similar to those seen in Georges Bank off New

D.B. Macurda, Jr.

England (Macurda, 1988).


The Zambesi River built an enormous delta into the area during the Late
Tertiary. The top of the delta plain (1) is illustrated in dip view in Figure 2
(Line 2). We see that this is notched by numerous submarine canyons (2) of
erosional origin that occur primarily at the shelf-slope break. Figure 3 (Line
1) is a strike line in the delta farther north along the same trend. Here, the
phenomena can be seen in a different perspective. The submarine canyons
form episodically, recording a lowering of base level. This phenomenon
happens simultaneously at different points along the margin associated with
the delta, over a distance of over 250 km. Is it allocyclic or autocyclic?
Figure 4 (Line 3) is a strike line farther north than the preceding lines,
away from the influence of deltaic sedimentation and possible delta
switching. Again, several levels of canyon formation are present, recording
recurrent lowering of base level. The oldest canyons are shown at (1).
Subsequent canyon development occurs at (2), with younger levels at (3)
and (4); the last two may be penecontemporaneous. Similar patterns
showing the recurrent formation of the erosional submarine canyons exist
for several hundred kilometers farther north along the Mozambique margin.
Thus, these are interpreted as representing allocyclic eventsthe recurrent
lowering of base level caused by the fluctuations of sea level due to
repetitive glaciations in the Miocene-Pleistocene.

REFERENCES CITED
Flores, G., 1973, The Cretaceous and Tertiary sedimentary basins of Mozambique
and Zululand, in G. Blant, ed., Sedimentary basins of the African coasts:
Paris Association of African Geological Surveys, p. 81-111.
Macurda, D.B., Jr., 1988, Contourites and volcanics, Georges Bank, New England:
this volume.
Martin, A.K., and C.J.H. Hartnady, 1986, Plate tectonic development of the
southwest Indian Oceana revised reconstruction of East Antarctica and
Africa: Journal of Geophysical Research, v. 91, no. B5, p. 4767-4786.
Mutter, John C., 1983, Structure within oceanic crust off the Norwegian margin, in
A.W. Bally, ed., Seismic expression of structural styles; a picture and work
atlas, volume 2: AAPG Studies in Geology 15, p. 2.2.3-36 to 2.2.3-38.
Mutter, J.C., M. Talwani, and P.L. Stoffa, 1982, Origin of seaward-dipping
reflectors in oceanic crust off the Norwegian margin by subaerial seafloor
spreading: Geology, v. 10, no. 7, p. 353-357.

Submarine canyons, contourites, and volcanics, Mozambique

265

RELATIVE SEA LEVEL CHANGES AND DEPOSITIONAL MODES


OF THE SHELF AND THE DEEP-SEA FAN OF THE INDUS

C. RAVENNE
Institut Francais du Petrole (IFP)
Rueil-Malmaison, France
F. COUMES
Societe Nationale Elf Aquitaine (SNEA(P))
Paris, France
and
J. P. ESTEVE
Compagnie Francaise des Petroles (CFP)
Paris, France
We show two profiles obtained during a Comite dEtudes Petrolies
Marines (CEPM) survey in which about 3000 km of seismic profiles were
obtained (see Index map, Figure 1). We focus on the Neogene sedimentary
fill, because during the Neogene the main canyon systems were incised on the
Indus shelf and slope, leading to the formation of channels and levees in the basin.
The canyon systems are analyzed on slope parallel profiles located near
the shelf edge. The channel systems are shown on a profile located at the
foot of the slope.
On the shelf (Profile Indus 10 (1) and (2)) the Neogene series can be
subdivided into two main seismic facies units: (1) as shelf facies unit,
visible on the edges of the profiles Indus 10 (1) and (2), and (2) a canyon
fill seismic facies unit that is cut into the shelf unit and that is visible in the
central portions of our figure (units A, B, and C).
The main characteristics of these two facies units are given on Table 1.
Within the Neogene we see a combination of shelf and canyon-fill units.
Profile Indus 10 (2) shows where the canyon-fill unit is subdivided into four parts
(A,B[1],B[2], and C) corresponding to superposed erosional and infilling phases.
Let us describe the process relating series 3 and 4 (Profile Indus 10 (2))
to canyon C. Erosion cuts into the shelf series 3, creating a canyon that then

C. Ravenne, F. Coumes and J.P. Esteve

is filled during the deposition of series 4, where we observe a clear genetic


relation between the shelf sediments of series 4 and the infilling of the
lower part of the canyon.
Beyond the present slope (Profile Indus 12), the Neogene sedimentary
series consists of a continuous package of stacked channels. The
characteristic seismic facies unit is the channel with its two parts, the
channel itself and the sedimentary levees.
The seismic characteristics of the channel show that it comprises two
parts. One part is a practically blind zone. The other part contains very
discontinuous reflectors, which are abruptly interrupted by high-tomoderate-amplitude reflectors of generally uniform frequency (high to
medium). In this part of the channel, the reflector configuration shows sets
of small, thick, parallel reflectors, grouped subhorizontally in aggregates
located at random within the transport zones, essentially in the axial zone.
In the levees bordering the channel, however, the reflectors are very
continuous, display a high- to moderate-amplitude rising progressively
toward the outer part of the levee, are of variable frequency, and are grouped
in convergent units toward the outer part of the levee.
The upper part of the channel-levee unit generally is covered by a
draping seismic facies unit commonly consisting of one or (sometimes) two
very continuous reflectors.
The lower part of the channel frequently corresponds to an erosional
truncation. The present Indus canyon is up to 10 km wide and more than 1
km deep on the outer margin of the shelf. The fossil canyons commonly
reach 10 km in width and may be over 15 km wide (canyon B2 in Profile
Indus 10 (2)). The channel-levee systems are just as large, and the width of
one complete system commonly exceeds 30 km. The channels may be up to
10 km wide, and each of the levees may be more than 50 km wide. The
maximum height of the levees is about 1 km.
Profile Indus 10-1 also shows the stepwise migration, from the
northwest to the southeast, of the Indus paleocanyon in its position, C, up to
its present position. The southeast flank displays a gentler slope affected by
many slides. The channel margins, between the upper parts of the levees,
also are affected by many slides (listric faults) that are not shown to avoid
complicating the interpretation.
Profile Indus 10 (2) shows the vertical sequence of depositional and
erosional phases.

Table 1. Main characteristics of the platform seismic facies units.


Shelf

Canyon

Continuity

generally continuous

very discontinuous short segments,


except at top and laterally

Amplitude

uniform along one reflection high,


moderate, or low

variable

Cycle breadth

varies like the amplitude

External form

sheet

Reflection geometry
at boundaries
Reflection geometry
of the sequence

concordant at the top and at the base


along depositional strike
concordant with gentle toplap at the
top and concordant with gentle
downlap at the base along
depositional dip

Reflection geometry
of the units

similar to the reflection geometry at


boundaries of the sequence, except
erosional truncation at contact with
canyon unit

Principal internal
configuration

gently oblique to sigmoid, parallel to commonly chaotic; parallel to


depositional dip; parallel along
divergent in the upper central part,
depositional strike
oblique to sigmoid near the edges

Lateral relations
of the sequence

may grade coastward to low


continuity and variable amplitude unit
(Indus Plain); commonly grade seaward
to progradational unit

channel fill

discordant at the base, concordant to


gentle toplap at the top, slight
general onlap on the edges

The Neogene (dating from well data indicates late Miocene to PlioceneQuaternary) began with the deposition of series 1. It was followed by the
erosion-transport of the first canyon, A (the term erosion-transport applies
to the dominant process, which began with erosion to excavate the canyon,
with simultaneous transport of erosion products seaward). The canyon then
acted as a drain, transporting toward the basin the terrigenous material
conveyed by the Indus River and material resulting from the erosion of the
coastal parts of the shelf. Lateral accretion and canyon-bed deposition may
have occurred simultaneously, but are of minor importance. Series 2a then
filled canyon A, followed by the erosion-transport of the second canyon,

Figure 1. Index showing location of seismic profiles Indus 10 and Indus


12, across the depositional fan of the Indus River.

Depositional modes of the shelf and deep sea fan of the Indus

270

B1. Series 2b was then deposited and eroded by canyon B2. Canyon B2
filled with the deposition of series 3, which was itself eroded subsequently
by canyon C, which in turn filled with the deposition of series 4. Finally,
series 5 then covered the shelf uniformly, and only the present canyon of the
Indus remains active.
The development of shelf seismic facies units (series 1, 2, 3, 4, and 5)
required a certain relief (relative elevation above sea level). The canyons
appear to have been formed at relatively low sea levels according to the
most general hypothesis, postulated among others by Vail et al. (1977).
Other data enabled us to determine that erosion became increasingly
dominant toward the coast, only allowing rare outliers of the previous
shelves to subsist. This suggests that toward the coast, erosion was
subaerial. This explanation, based on water depth variations on the shelf,
appears to be supported by the number of erosion periods, associated with
low water depth, and with a similar number of periods of levee construction
in the basin.
In the case of the present Indus [Profile Indus 10 (1)], the canyon cuts
increasingly deeper into the platform in a seaward direction. If a relative
drop in sea level affects this zone, erosion occurs and all the sediments of
the zone near the coast are reworked. A relative rise of sea level, succeeding
this phase, again allows shelf sedimentation and submerges the canyon
(which therefore only appears to begin at the middle of the shelf).
The formation of the upper Miocene to Pliocene-Quaternary unit below
the present shelf of the Indus basin can be summarized as follows: (1) Shelf
deposition occurred during periods of relative rise of sea level; and (2)
Complex erosion-sedimentation (braided stream type) took place near the
coast, with the formation of canyons seaward during a period of relative
drop in sea level.
During periods of relative lowering of the sea level, the material from
river deposition and shelf stripping led to the formation of the Indus Deep
Sea Fan, with its channel-levee system near the slope base.
During periods of relative rise of sea level, the river deposits were
trapped on the shelf. In the basin, only draping strata of the channel-levee
systems were deposited (pelagic sediments, fine, distal parts of turbidites,
and nepheloid layers).
To conclude, we postulate that the construction of the shelf during the
late Miocene to the Pliocene-Quaternary period took place during four
major periods of relative elevation of sea level, interrupted by three periods
of a relative drop in sea level. These three drops correspond to the periods
of canyon formation, culminating in the construction of the channel-levee
systems. Periods of relative elevation of sea level are reflected in the basin
by the deposition of draping strata.

C. Ravenne, F. Coumes and J.P. Esteve

ACKNOWLEDGMENTS
The authors wish to thank the CEPM for allowing the shooting and
analysis of these profiles.
Their sincere thanks also go to Messrs. M. Larere and E.L. Nico of
SNEA(P) and M. Becquey (IFP) for their active participation in this study.

REFERENCE CITED
Vail, P.R., et al., 1977, Seismic stratigraphy and global changes of sea level, in C.E.
Payton, ed., Seismic stratigraphy, application to hydrocarbon exploration:
AAPG Memoir 26, p. 49-212.

GEOPHYSICAL PARAMETERS
Energy source:
Stacking multiplicity:
Number of channels recorded:
Interval between input channels:
Minimum offset distance:
Maximum offset distance:

Vaporchoc (CGG)
24
24
100 m
290 m
2590 m

Depositional modes of the shelf and deep sea fan of the Indus

271

You might also like