You are on page 1of 552

Preface

T his memoir addresses the geology and dynamics of the principal petroleum basins of South
America from Venezuela to Argentina. Our understanding of the way these sedimentary
basins formed has advanced immensely in the last decade. This reflects improved data quality, a
different way of looking at the data, a new integrated approach to basin analysis, and a new genera-
tion of geologists. A wide variety of papers reviews the tectonic framework, comparative structural
styles, sequence stratigraphy, and the way these are interrelated. These basins were a long time in
the making. Most have evolved by repeated reactivation of preexisting fabrics. The structural devel-
opment, its timing, and the way sedimentation responded to this activity is of paramount impor-
tance to oil and gas exploration. Intentionally we do not review in any detail the Mesozoic basins of
the Atlantic margin because a subject so vast would require space of its own.
We are grateful to the following people for their generous assistance in many ways, for being
there when we needed them: Peter Aukes, Hugo Belotti, Nora Cesaretti, Miguel Cirbián, Al
Ferworn, Silvia González, Gary Howell, Bob Meneley, Luis Spalletti and Graciela Suárez Marzal de
Spalletti, Anne Thomas, Gustavo Vergani, and members of the AAPG staff. We thank Edgar Ortiz
and Jeremy Tankard for the paintings prepared especially for this volume. Copyediting and layout
were done by Kathy Walker. We are also grateful to Perez Companc for its support.
All of the manuscripts were reviewed externally. We would like to thank the following for their
critical reviews:

J. Allan D. A. Lehto
R. B. Allen J. Letouzey
R. W. Allmendinger S. Liu
H. Arbe O. López Gamundi
C. D. Arregui O. López Paulsen
P. G. Aukes D. Loureiro
H. R. Balkwill J. D. Lowell
C. Barcat J. A. Maloney
H. Belotti P. Mann
K. T. Biddle R. Manoni
A. Boll J. R. McLean
G. C. Bond R. A. Meneley
A. J. Boucot E. J. Milani
R. L. Brown M. Mozetic
K. R. Butler A. C. Newson
R. Caminos H. Passalacqua
M. A. Cooper M. J. Perkins
M. C. Covey J. L. Pindell
L. Dalla Salda V. Ploszkiewicz
M. J. de Wit V. A. Ramos
J. F. Dunn J. Reynolds
A. Edwards G. Rodrigue
R. N. Erlich D. Roeder
N. Eyles F. Roure
R. D. Forsythe D. B. Rowley
J. Gallagher F. Schein
C. F. Garrasino D. W. Scholl
S. Greer T. Sempere
R. H. Groshong B. M. Sheffels
N. E. Haimila A. E. Slingsby
P. E. Isaacson L. R. Snowdon
M. P. A. Jackson C. R. Tippett
W. R. Jamison M. A. Uliana
W. A. M. Jenkins J. Utting
P. B. Jones B. van Hoorn
T. E. Jordan G. Vergani
J. W. Kerr J. N. J. Visser
E. Kozlowski C. Vistalli
C. Knutson H. J. Welsink
G. E. Kronman H. J. White
D. Larue P. A. Ziegler
L. Legarreta N. Zilli

This memoir was stimulated by a meeting on Gondwana geology held in Santa Cruz, Bolivia, in
August 1992. Financial support was provided by Yacimientos Petroliferos Fiscales Bolivianos,
Phillips Petroleum, Tesoro Bolivia Petroleum, Perez Companc, Mobil Boliviana Petroleum, Chevron
International (Bolivia), Exxon Company, and Texaco Bolivia.

Anthony Tankard
Ramiro Suárez Soruco
Herman Welsink
About the Editors

ANTHONY TANKARD is a principal R AMIRO S UAREZ S ORUCO is a H ERMAN W ELSINK is a senior


of Tankard Enterprises, a senior technical specialist at geologist in the exploration
Calgary-based company. A native Yacimientos Petrolíferos Fiscales department of Compañía Naviera
of South Africa, he received B.Sc. Bolivianos, the Bolivian national Perez Companc S.A.C.F.I.M.F.A. in
(Honours) and Ph.D. degrees oil company. Ramiro was born in Neuquén, Argentina. He was
from Natal University and Cochabamba in the Andes. He born in Haarlem, The Nether-
Rhodes University, respectively. completed his undergraduate lands. He graduated with a
Since 1970, he worked in research training at the Universidad Doctoraal in geology from the
institutions in South Africa and Nacional de La Plata and his University of Utrecht; this work
the United States before entering Doctorate in geology at the included a thesis on an Eocene
the Canadian petroleum industry Universidad de Buenos Aires, basin in the Pyrenees. Since 1981,
in 1981. Tony established Tankard both in Argentina. Since 1968, his Herman has worked in the oil
Enterprises in 1992. His work career has spanned GEOBOL and industry, where he has been
includes basin evolution, stratig- YPFB as paleontologist, biostratig- involved in exploration princi-
raphy, the way sedimentation rapher, and technical services pally of extensional basins in
responds to structural develop- manager. Ramiro’s research Canada, Bolivia, and Argentina.
ment, and the habitat of hydro- interests have focused on inverte- His research interests and publi-
carbon accumulation. He is brate paleontology and Paleozoic cations center on the formation
actively involved in AAPG affairs biostratigraphy and paleogeog- and deformation of sedimentary
and has published numerous raphy, subjects in which he has basins, structural-stratigraphic
technical papers in addition to co- published many papers. Through linkage, and their consequences
editing or co-authoring several this work, he has been involved in for hydrocarbon accumulation.
volumes. several national and international
committees, including the IUGS
Subcommission on Devonian
Stratigraphy and the Academia
Nacional de Ciencias de Bolivia.
About the Artists

E dgar Ortiz was born in Villa Dolores in the Province of Córdoba, Argentina. He studied
drawing and painting at the Manuel Belgrano School of Fine Arts. Today Ortiz is widely
regarded as one of Argentina’s outstanding watercolor artists, for which he has received numerous
awards. His work is inspired by the Argentinian landscapes and captures their changing seasons
and moods. His paintings are in numerous galleries as well as private collections in Argentina,
Brazil, Venezuela, Mexico, Canada, United States, Italy, and The Netherlands. Edgar now lives in La
Plata.

J eremy Tankard was born in Cape Town, South Africa. He is now resident in Canada, where he
studied drawing and printmaking at the Alberta College of Art. He is principally a figure and
portrait artist and works in most mediums. Jeremy lives in Calgary.
Petroleum Basins of
South America

edited by

A. J. Tankard

R. Suárez Soruco

H. J. Welsink

AAPG Memoir 62

Published jointly by

The American Association of Petroleum Geologists

Yacimientos Petrolíferos Fiscales Bolivianos

Academia Nacional de Ciencias de Bolivia

Printed in the U.S.A.


Copyright © 1995
The American Association of Petroleum Geologists
All Rights Reserved
Published 1995

ISBN: 0-89181-341-1

AAPG grants permission for a single photocopy of an item from this publication for personal use.
Authorization for additional copies of items from this publication for personal or internal use is
granted by AAPG provided that the base fee of $3.00 per copy is paid directly to the Copyright
Clearance Center, 222 Rosewood Drive, Danvers, Massachusetts 01923. Fees are subject to change.
Any form of electronic or digital scanning or other digital transformation of portions of this publica-
tion into computer-readable and/or transmittable form for personal or corporate use requires
special permission from, and is subject to fee charges by, the AAPG.

Association Editor: Kevin T. Biddle


Science Director: Richard Steinmetz
Publications Manager: Kenneth M. Wolgemuth
Special Projects Editor: Anne H. Thomas
Production and Layout: Kathy and Dana Walker, Editorial Technologies

THE AMERICAN ASSOCIATION OF PETROLEUM GEOLOGISTS


(AAPG) DOES NOT ENDORSE OR RECOMMEND ANY PRODUCTS
OR SERVICES THAT MAY BE CITED, USED, OR DISCUSSED IN
AAPG PUBLICATIONS OR IN PRESENTATIONS AT EVENTS ASSO-
CIATED WITH AAPG.
About the Editors

ANTHONY TANKARD is a principal R AMIRO S UAREZ S ORUCO is a H ERMAN W ELSINK is a senior


of Tankard Enterprises, a senior technical specialist at geologist in the exploration
Calgary-based company. A native Yacimientos Petrolíferos Fiscales department of Compañía Naviera
of South Africa, he received B.Sc. Bolivianos, the Bolivian national Perez Companc S.A.C.F.I.M.F.A. in
(Honours) and Ph.D. degrees oil company. Ramiro was born in Neuquén, Argentina. He was
from Natal University and Cochabamba in the Andes. He born in Haarlem, The Nether-
Rhodes University, respectively. completed his undergraduate lands. He graduated with a
Since 1970, he worked in research training at the Universidad Doctoraal in geology from the
institutions in South Africa and Nacional de La Plata and his University of Utrecht; this work
the United States before entering Doctorate in geology at the included a thesis on an Eocene
the Canadian petroleum industry Universidad de Buenos Aires, basin in the Pyrenees. Since 1981,
in 1981. Tony established Tankard both in Argentina. Since 1968, his Herman has worked in the oil
Enterprises in 1992. His work career has spanned GEOBOL and industry, where he has been
includes basin evolution, stratig- YPFB as paleontologist, biostratig- involved in exploration princi-
raphy, the way sedimentation rapher, and technical services pally of extensional basins in
responds to structural develop- manager. Ramiro’s research Canada, Bolivia, and Argentina.
ment, and the habitat of hydro- interests have focused on inverte- His research interests and publi-
carbon accumulation. He is brate paleontology and Paleozoic cations center on the formation
actively involved in AAPG affairs biostratigraphy and paleogeog- and deformation of sedimentary
and has published numerous raphy, subjects in which he has basins, structural-stratigraphic
technical papers in addition to co- published many papers. Through linkage, and their consequences
editing or co-authoring several this work, he has been involved in for hydrocarbon accumulation.
volumes. several national and international
committees, including the IUGS
Subcommission on Devonian
Stratigraphy and the Academia
Nacional de Ciencias de Bolivia.

iii
JUAN KEIDEL (1877–1954)
Geologist, teacher, and explorer.

Jeremy Tankard, 1994, watercolor, gouache, ink and


charcoal, 38 × 41 cm
Dedication

B efore plate tectonics, there was continental drift, a concept pioneered largely by Alfred
Wegener. He showed that the geometric reconstruction of the supercontinents was more than
the fortuitous parallelism of the coasts on either side of the Atlantic. He also explained in an elegant
way the distribution of mountain belts, stratigraphy, and middle Paleozoic ice age deposits.
The reaction to these ideas was lukewarm at best in Europe and decidedly hostile in North
America. The geologists of southern Africa and South America were, however, more enthusiastic.
Prominent among these were Alex du Toit of South Africa and Juan Keidel of Argentina.
Juan Keidel provided some of the supporting evidence for the contiguity of Africa and South
America within Paleozoic Gondwana. His contributions are acknowledged by Wegener in his
classic book Die Entstehung der Kontinente und Ozeane (1915). Keidel demonstrated the relationships
between the Carboniferous–Permian Sauce Grande glacial deposits of Argentina and the Dwyka
tillites of South Africa, as well as the nature of their encapsulating stratigraphies. Only in a recon-
structed Gondwana did these glaciers have a rational distribution. The apparent continuity of the
Cape fold belt and the Sierra de la Ventana was noted. In his seminal paper of 1921, Keidel recog-
nized how widespread this Permian–Triassic deformation really was, forming a series of cordilleras
from Ventana to the Andean foothills. To describe these ancient cordilleras, he coined the name
Gondwanides. These interpretations were supported by field work on the Ventana System in the
province of Buenos Aires and by surveys in the western and northern Precordillera of Mendoza and
San Juan. This was also the start of his association with Alex du Toit, who visited Argentina in the
early 1920s.
Field expeditions took Keidel to the Andes, Patagonia, and Neuquén. In the Neuquén basin, he
was involved with the first oil well.
Juan Keidel was born in Gross Stoeckheim, Germany, in 1877. He studied at the Institute of
Mining in Berlin and completed a doctorate at Freiburg under Professor Steinmann. Dr. Keidel was
appointed the first head of what was to become the Geological Survey of Argentina. Subsequently,
he taught geology at the University of Buenos Aires. Keidel published more than 50 papers in an
illustrious career.
We dedicate this book, Petroleum Basins of South America, to the memory of Juan Keidel.

v
Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .xi

Regional Setting
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana
During the Phanerozoic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5
A. J. Tankard, M. A. Uliana, H. J. Welsink, V. A. Ramos, M. Turic, A. B. França, E. J.
Milani, B. B. de Brito Neves, N. Eyles, J. Skarmeta, H. Santa Ana, F. Wiens, M. Cirbián,
O. López P., G. J. B. Germs, M. J. De Wit, T. Machacha, and R. McG. Miller
Oil and Gas Discoveries and Basin Resource Predictions in Latin America . . . . . . . . . . . . . . . . . . .53
G. E. Kronman, S. W. Rushworth, K. Jagiello, and A. Aleman
Petroleum Basins of Southern South America: An Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .63
C. M. Urien, J. J. Zambrano, and M. R. Yrigoyen
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana . . . . . . . . . . . . . . . . .79
K. E. Williams
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on
Hydrocarbon Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .101
J. L. Pindell and K. D. Tabbutt
Phanerozoic Correlation in Southern South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .129
A. B. França, E. J. Milani, R. L. Schneider, O. López P., J. López M., R. Suárez S., H. Santa
Ana, F. Wiens, O. Ferreiro, E. A. Rossello, H. A. Bianucci, R. F. A. Flores, M. C. Vistalli,
F. Fernandez-Seveso, R. P. Fuenzalida, and N. Muñoz

Paleozoic Basins
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and
Central South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .165
N. Eyles, G. Gonzalez Bonorino, A. B. França, C. H. Eyles, and O. López Paulsen
Phanerozoic Tectonics and Sedimentation in the Chaco Basin of Paraguay,
with Comments on Hydrocarbon Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .185
F. Wiens
Phanerozoic Evolution of Bolivia and Adjacent Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .207
T. Sempere
Evidence for a Middle–Late Paleozoic Foreland Basin and Significant Paleolatitudinal Shift,
Central Andes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .231
P. E. Isaacson and E. Díaz Martínez
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina . . . . . . . . . . . . .251
D. Starck
Tectonic Evolution of the Andes of Northern Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .269
R. Mon and J. A. Salfity
Tectonics and Stratigraphy of the Late Paleozoic Paganzo Basin of Western Argentina and its
Regional Implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .285
F. Fernandez-Seveso and A. J. Tankard

vii
viii Contents

Mesozoic Rifts
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia . . . . . . . . . . . . . . . . .305
H. J. Welsink, E. Martinez, O. Aranibar, and J. Jarandilla
Geometry and Seismic Expression of the Cretaceous Salta Rift System,
Northwestern Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .325
A. H. Comínguez and V. A. Ramos
Cretaceous Rifting, Alluvial Fan Sedimentation, and Neogene Inversion,
Southern Sierras Pampeanas, Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .341
C. J. Schmidt, R. A. Astini, C. H. Costa, C. E. Gardini, and P. E. Kraemer
Structural Inversion and Oil Occurrence in the Cuyo Basin of Argentina . . . . . . . . . . . . . . . . . . . .359
D. Dellapé and A. Hegedus
Inversion of the Mesozoic Neuquén Rift in the Malargüe Fold and Thrust Belt,
Mendoza, Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .369
R. Manceda and D. Figueroa
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina . . . . . . . . . . . . . . . . . .383
G. D. Vergani, A. J. Tankard, H. J. Belotti, and H. J. Welsink
Hydrocarbon Accumulation in an Inverted Segment of the Andean Foreland:
San Bernardo Belt, Central Patagonia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .403
G. O. Peroni, A. G. Hegedus, J. Cerdan, L. Legarreta, M. A. Uliana, and G. Laffitte

Andean Basins
Petroleum Geology of the Sub-Andean Basins of Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .423
J. M. P. Mathalone and M. Montoya R.
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone . . . . . . . . . . . . . . . . .445
P. Baby, I. Moretti, B. Guillier, R. Limachi, E. Mendez, J. Oller, and M. Specht
Structural Geology of Sub-Andean Fold and Thrust Belt in Northwestern Bolivia . . . . . . . . . . . .459
D. Roeder and R. L. Chamberlain
Andean and Pre-Andean Deformation, Boomerang Hills Area, Bolivia . . . . . . . . . . . . . . . . . . . . .481
H. J. Welsink, A. Franco M., and C. Oviedo G.
Devonian–Carboniferous Stratigraphy in the Madre de Dios Basin, Bolivia:
Pando X-1 and Manuripi X-1 Wells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .501
P. E. Isaacson, B. A. Palmer, B. L. Mamet, J. C. Cooke, and D. E. Sanders
Is the Bend in the Bolivian Andes an Orocline? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .511
B. M. Sheffels
Structural Styles and Hydrocarbon Potential of the Sub-Andean Thrust Belt
of Southern Bolivia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .523
J. F. Dunn, K. G. Hartshorn, and P. W. Hartshorn
Structural Styles and Petroleum Occurrence in the Sub-Andean Fold and
Thrust Belt of Northern Argentina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .545
H. J. Belotti, L. L. Saccavino, and G. A. Schachner
Contents ix

Northern South America


Northern Part of Oriente Basin, Ecuador: Reflection Seismic Expression
of Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .559
H. R. Balkwill, G. Rodrigue, F. I. Paredes, and J. P. Almeida
Reservoir Characterization of the Hollin and Napo Formations,
Western Oriente Basin, Ecuador . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .573
H. J. White, R. A. Skopec, F. A. Ramirez, J. A. Rodas, and G. Bonilla
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru:
Sedimentologic and Tectonic Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .597
R. Marocco, A. Lavenu, and R. Baudino
Basin Development in an Accretionary, Oceanic-Floored Fore-Arc Setting:
Southern Coastal Ecuador During Late Cretaceous–Late Eocene Time . . . . . . . . . . . . . . . . . . .615
É. Jaillard, M. Ordoñez, S. Benitez, G. Berrones, N. Jiménez, G. Montenegro,
and I. Zambrano
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution . . . . . . . . . . . . . . . . . . . . . . .633
D. Roeder and R. L. Chamberlain
Geodynamic Evolution of the Eastern Andes, Colombia: An Alternative Hypothesis . . . . . . . .647
P. B. Jones
Basin Development and Tectonic History of the Llanos Basin, Colombia . . . . . . . . . . . . . . . . . . . .659
M. A. Cooper, F. T. Addison, R. Alvarez, A. B. Hayward, S. Howe, A. J. Pulham,
and A. Taborda
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges . . . . . . . . . . . . . .667
H. Passalacqua, F. Fernandez, Y. Gou, and F. Roure
Stratigraphic Synthesis of Western Venezuela . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .681
F. Parnaud, Y. Gou, J.-C. Pascual, M. A. Capello, I. Truskowski, and H. Passalacqua
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela . . . . . . . . . . . . . . . . . . . . . . . .699
J. Lugo and P. Mann
Two-Dimensional Computer Modeling of Oil Generation and Migration in a
Transect of the Eastern Venezuela Basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .727
O. Gallango and F. Parnaud
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin . . . . . . . . . . . . . . . . . . . .741
F. Parnaud, Y. Gou, J.-C. Pascual, I. Truskowski, O. Gallango, H. Passalacqua, and F. Roure
Cenozoic Sedimentation and Tectonics of the Southwestern Caribbean Pull-Apart
Basin, Venezuela and Colombia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .757
C. E. Macellari

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .781
Preface

T his memoir addresses the geology and dynamics of the principal petroleum basins of South
America from Venezuela to Argentina. Our understanding of the way these sedimentary
basins formed has advanced immensely in the last decade. This reflects improved data quality, a
different way of looking at the data, a new integrated approach to basin analysis, and a new genera-
tion of geologists. A wide variety of papers reviews the tectonic framework, comparative structural
styles, sequence stratigraphy, and the way these are interrelated. These basins were a long time in
the making. Most have evolved by repeated reactivation of preexisting fabrics. The structural devel-
opment, its timing, and the way sedimentation responded to this activity is of paramount impor-
tance to oil and gas exploration. Intentionally we do not review in any detail the Mesozoic basins of
the Atlantic margin because a subject so vast would require space of its own.
We are grateful to the following people for their generous assistance in many ways, for being
there when we needed them: Peter Aukes, Hugo Belotti, Nora Cesaretti, Miguel Cirbián, Al
Ferworn, Silvia González, Gary Howell, Bob Meneley, Luis Spalletti and Graciela Suárez Marzal de
Spalletti, Anne Thomas, Gustavo Vergani, and members of the AAPG staff. We thank Edgar Ortiz
and Jeremy Tankard for the paintings prepared especially for this volume. Copyediting and layout
were done by Kathy Walker. We are also grateful to Perez Companc for its support.

xi
xii Preface

All of the manuscripts were reviewed externally. We would like to thank the following for their
critical reviews:

J. Allan D. A. Lehto
R. B. Allen J. Letouzey
R. W. Allmendinger S. Liu
H. Arbe O. López Gamundi
C. D. Arregui O. López Paulsen
P. G. Aukes D. Loureiro
H. R. Balkwill J. D. Lowell
C. Barcat J. A. Maloney
H. Belotti P. Mann
K. T. Biddle R. Manoni
A. Boll J. R. McLean
G. C. Bond R. A. Meneley
A. J. Boucot E. J. Milani
R. L. Brown M. Mozetic
K. R. Butler A. C. Newson
R. Caminos H. Passalacqua
M. A. Cooper M. J. Perkins
M. C. Covey J. L. Pindell
L. Dalla Salda V. Ploszkiewicz
M. J. de Wit V. A. Ramos
J. F. Dunn J. Reynolds
A. Edwards G. Rodrigue
R. N. Erlich D. Roeder
N. Eyles F. Roure
R. D. Forsythe D. B. Rowley
J. Gallagher F. Schein
C. F. Garrasino D. W. Scholl
S. Greer T. Sempere
R. H. Groshong B. M. Sheffels
N. E. Haimila A. E. Slingsby
P. E. Isaacson L. R. Snowdon
M. P. A. Jackson C. R. Tippett
W. R. Jamison M. A. Uliana
W. A. M. Jenkins J. Utting
P. B. Jones B. van Hoorn
T. E. Jordan G. Vergani
J. W. Kerr J. N. J. Visser
E. Kozlowski C. Vistalli
C. Knutson H. J. Welsink
G. E. Kronman H. J. White
D. Larue P. A. Ziegler
L. Legarreta N. Zilli

This memoir was stimulated by a meeting on Gondwana geology held in Santa Cruz, Bolivia, in
August 1992. Financial support was provided by Yacimientos Petroliferos Fiscales Bolivianos,
Phillips Petroleum, Tesoro Bolivia Petroleum, Perez Companc, Mobil Boliviana Petroleum, Chevron
International (Bolivia), Exxon Company, and Texaco Bolivia.

Anthony Tankard
Ramiro Suárez Soruco
Herman Welsink
About the Artists

E dgar Ortiz was born in Villa Dolores in the Province of Córdoba, Argentina. He studied
drawing and painting at the Manuel Belgrano School of Fine Arts. Today Ortiz is widely
regarded as one of Argentina’s outstanding watercolor artists, for which he has received numerous
awards. His work is inspired by the Argentinian landscapes and captures their changing seasons
and moods. His paintings are in numerous galleries as well as private collections in Argentina,
Brazil, Venezuela, Mexico, Canada, United States, Italy, and The Netherlands. Edgar now lives in La
Plata.

J eremy Tankard was born in Cape Town, South Africa. He is now resident in Canada, where he
studied drawing and printmaking at the Alberta College of Art. He is principally a figure and
portrait artist and works in most mediums. Jeremy lives in Calgary.

xiii
Petroleum Basins of
South America
Regional Setting

MACIZO DEL FITZROY or El Chaltén, Santa Cruz province, Argentina.


These granitic needles (3405 m) are part of the Paleocene–Eocene Serie
Andesítica, a string of plutons in the Andes Australes.

Edgar Ortiz, 1994, watercolor, 30 × 23 cm


Oil and Gas Discoveries and Basin Resource
Predictions in Latin America

George E. Kronman

Sandra W. Rushworth

Keith Jagiello

Antenor Aleman
Amoco Production Company
Houston, Texas, U.S.A.

Abstract

M ore than 4500 wildcat wells were drilled from 1980 to 1990 in South America. Approximately 355 of
these resulted in hydrocarbon discoveries. An estimated 12% of the discoveries contain reserves
greater than 100 MMBO. Several of the larger finds (>500 MMBO), such as Cusiana (Colombia),
Furrial–Musipan (Venezuela), Caño Limón (Colombia), and Marlim (Brazil), are important among the giant
fields found worldwide since 1980. Most of the larger discoveries were made by national oil companies in
Venezuela, Mexico, and Brazil. The probability of finding large oil fields (>500 MMBO) is greatest in the
Campos, Llanos, Reforma-Campeche, Maracaibo, and Maturin basins. Smaller, but still significant fields
(50–250 MMBO) may still be found in the Neuquén, San Jorge, Austral, Tarija, Marañon-Napo-Putumayo,
Magdalena, and Tampico-Misantla basins. More than 170 BBO of proven reserves have been found in the
highest potential Latin American basins. Undiscovered oil resources of 40–80 BBO are estimated to remain in
the same group based on historical field size data and current geologic knowledge. Frontier and emerging
basins may also contain significant resources, but limited data makes it difficult to estimate their undiscovered
potential.

Resumen

M as de 4500 pozos de exploración se han perforado desde 1980 hasta 1990 en América Latina. Aproxi-
madamente 355 de estos resultaron en descubrimiento de hidrocarburo. Se estima que un doce por
ciento (12%) de los descubrimientos contienen reservas mayores que 100 MMBO. Varios de los hallazgos
mayores (>500 MMBO), tales como Cusiana (Colombia), Furrial/Musipan (Venezuela), Caño Limón
(Colombia), y Marlim (Brazil) clasifican entre los campos gigantes mas importantes descubiertos a nivel
mundial desde 1980. La mayoría de los descubrimientos mas importantes fueron hechos por compañias
nacionales en Venezuela, México y Brazil. La probabilidad de encontrar nuevos yacimientos gigantes de
petróleo (>500 MMBO) es mayor en las cuencas Campos, Llanos, Reforma-Campeche, Maracaibo y Maturin.
Yacimientos menores, pero aun importantes (50–250 MMBO) pueden todavía ser descubiertos en las cuencas
de Neuquén, San Jorge, Austral, Tarijas, Marañon-Oriente-Putumayo, Magdalena y Tampico-Misantla. Sobre
170 BBO de reservas probadas se han encontrado en el mismo grupo de cuencas. Se estima que los recursos de
hidrocarburo por descubrirse en las cuencas Latinoamericanas de mayor potencial, basado en los datos de la
historia del tamaño del campo y el conocimiento geológico actual, son de 40 a 80 MMBO. Las cuencas
nacientes y áreas exploratorias de frontera también pueden contener recursos significativos, pero los escasos
datos disponibles hacen difícil estimar su potencial.

Kronman, G. E., Rushworth, S. W., Jagiello, K., and Aleman, A., 1995, Oil and gas discov- 53
eries and basin resource predictions in Latin America, in A. J. Tankard, R. Suárez S., and
H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 53–61.
54 Kronman et al.

INTRODUCTION
Latin America has the second largest oil reserves in
the world, after the Middle East (Figure 1). Increased
democratization, restructuring toward free market
economic systems, privatization, and growing confi-
dence in the potential for Latin America are attracting
significant interest in the area. As a result, many of the
world’s multinational petroleum companies are increas-
ingly pursuing exploration and production opportunities
in Latin America (Kronman et al., 1993).
This paper presents exploration wildcat well success
rates, estimated field discovery sizes, and competitor
information from 1980 to 1992 throughout Latin
America. We have used this information to estimate
remaining resources, as well as probabilities of finding
these resources in the larger producing Latin American
basins. The information presented here is based largely
Figure 1—Crude oil resources greater than or equal to
on an internal study done by Amoco on Latin America. 2 BBO recoverable by country as of January 1, 1991.
Data used for this study are complete through the end of (Source of information: Oil and Gas Journal Energy
1990. In some cases, our data base after 1990 is incom- Database.)
plete or unverified and thus is not used. For the purpose
of this paper, Latin America includes all countries from
Mexico to the southern tip of Latin America, including
the Caribbean nations (Figure 2).
The data used in this study are based on outside
resource estimates, Amoco’s proprietary database, and
estimates made by Amoco in-house experts in Latin
America. All field size data and estimates are recorded as
recoverable reserves. Gas reserves are converted to
barrels of oil equivalent (1 BOE = 6000 cf gas), and all
reserves are stated as barrels of oil or oil equivalent.

LATIN AMERICA FROM 1980 TO 1990


Wildcat Drilling
From 1980 to 1990, about 4522 wildcat wells were
drilled in ten principal oil-prone countries of Latin
America: Argentina, Bolivia, Brazil, Colombia, Ecuador,
Guatemala, Peru, Surinam, Trinidad, and Venezuela
(Figure 2). Mexican and Cuban wildcat well data were
unavailable. Approximately 82% of these wells were dry.
About 12% were considered “technical successes” to the
extent that they flowed some oil or gas on a production
test and reported an accumulation of less than 100
million barrels of oil equivalent (MMBOE). Many of
these wells were plugged and have questionable
economic viability. Another 4% of the wildcat wells led
to gas field discoveries. About 1% of the wildcats discov-
ered fields with resources between 100 and 500 million
barrels of oil (MMBO), and about 0.3% of the wildcat
wells led to the discovery of giant oil fields with reserves
greater than 500 MMBO.
When the percentages of wildcat success in Latin
America are compared with the rest of the world
(excluding the United States and Canada), there is a
strong correlation of results (Figure 3). The percentage of Figure 2—Map of Latin America showing major petroleum
wildcats in Latin America that discovered oil fields basins mentioned in this paper.
Oil and Gas Discoveries and Basin Resource Predictions in Latin America 55

(a)

(a)

(b)

Figure 4—Analysis of wildcat wells drilled in Latin America


for 1980–1990. (a) Total number of wildcat wells drilled in
10 Latin American countries, including a breakdown of
(b) “successful” wildcat wells and percent success by
country. The same 10 countries listed in Figure 3 are
included here. (b) Total number and percent success of
wildcat wells drilled in the same countries shown on an
Figure 3—(a) Wildcat well field discovery percentages for annual basis and as an average for 1980–1990. Note that
all wildcat wells drilled in 10 Latin American countries the largest number of wildcat wells were drilled in the early
(Argentina, Bolivia, Brazil, Colombia, Ecuador, Guatemala, 1980s, but the average success percentage of wildcat wells
Peru, Suriname, Trinidad, and Venezuela; data from Cuba increased from 1985 to 1988.
and Mexico not available). (b) Wildcat well field discovery
percentages for all wildcat wells drilled worldwide, (Figure 4a). Less than 100 wildcat wells each were drilled
excluding the United States and Canada, for 1980–1990.
in Ecuador, Peru, Bolivia, Trinidad, Guatemala, and
The field discovery percentages for Latin America are
similar to those worldwide. Surinam. No data were available for Mexico and Cuba.
The average “success ratio” for wildcat drilling during
this period was highest in Ecuador (about 60%), followed
between 100 and 500 MMBO and greater than 500 by Brazil (about 38%), Trinidad (about 32%), and Peru
MMBO is comparable to worldwide discoveries. The (about 30%) (Figure 4a).
only significant difference is that the percentage of gas During this 1980–1990 period in Latin America, the
fields discovered elsewhere in the world (7%) is higher largest number of wildcat wells were drilled from 1980 to
than in Latin America (4%). The reason for this difference 1982. Between 550 and 650 wells were drilled each year
is unknown, but it may reflect the general lack of an during these 3 years (Figure 4b). Subsequently, the
indigenous gas market or distribution system in parts of number of wildcat wells drilled declined to about 210 in
Latin America, as a result of which some gas discoveries 1990. The average number of wells drilled each year in
are unreported. Latin America over the total 10-year period was 396.
The largest number of wildcat wells drilled in Latin Wildcat success rates were about 25% in 1980, declining
America during the 1980–1990 period was in Brazil, to less than 20% in 1983, and rebounding to almost 30%
followed by Venezuela, Argentina, and Colombia in 1988 before falling again to 21% in 1990.
56 Kronman et al.

Field Discoveries
There were 355 reported oil field discoveries in Latin
America from 1980 to 1990. All occurred within the
countries previously mentioned, except for two discov-
eries in Chile and one in Panama (information was not
available for Cuba). Forty-four of the Latin American oil
field discoveries (12%) are estimated to contain resources
greater than 100 MMBO.
Most of the larger field discoveries (>100 MMBO)
were made in Venezuela (fifteen), followed by Brazil
(twelve), Mexico (seven), Colombia (five), Ecuador (two),
and one each in Peru, Trinidad, and Argentina (Figure
5a). All of the field discoveries in Venezuela, Brazil, and
Mexico were made by their national oil companies,
which have exclusive rights to the discoveries.
Most of the large field discoveries (>100 MMBO)
occurred in 1980 (six), 1988 (seven), and 1989 (ten).
Figure 5b shows that the annual number of large field
discoveries declined from six in 1980 to one in 1983,
remained relatively stable from 1984 to 1987 (three to
five), and increased to ten in 1989.
The largest oil and gas fields discovered in the
principal Latin American countries from 1980 to 1992 are
listed in Tables 1 and 2. Fields estimated to be greater
than 1 BBO include Marlim in Brazil, Furrial-Musipan in
Venezuela, Euch in Mexico, and Cusiana in Colombia.
Gas fields greater than 1 tcf were found in Peru, Brazil,
Colombia, Venezuela, and Argentina.

Competitor Analysis
Three groups of competitors have been identified in
Latin America. The national oil companies (NOCs) and
multinational oil companies (MNCs) are the major
competitors. The activities of the smaller independents
are more limited, but in general, they have become
increasingly active and successful. NOCs are separate,
identifiable businesses that are engaged in either the
upstream or downstream parts of the petroleum
industry whose interest is controlled by a national
government and that serve national interests (Kronman
and Smith, 1993). The larger NOCs in Latin America
include Petroleos de Venezuela (PDVSA) in Venezuela,
Petrobras in Brazil, Pemex in Mexico, Ecopetrol in
Colombia, and Petroecuador in Ecuador. Multinational
companies with a significant commitment and invest-
ment in Latin America include Occidental, Texaco, Elf,
Exxon, Royal Dutch/Shell, British Petroleum, Amoco,
Mobil, ARCO, Chevron, and Conoco.
During the 1980–1990 period, the Latin American
NOCs had 35 significant discoveries (greater than or
Figure 5—Major field discoveries for 1980–1990 greater
than 100 MMBO made (a) by country, (b) on an annual
equal to 100 MMBO), which is four times greater than
basis, and (c) by company in Latin America. These discov- the 8 significant oil discoveries made by the multina-
eries are based on various external sources and the tional oil company group (Figure 5c). Two NOCs,
Amoco Production Company Resource Assessment PDVSA (Venezuela) and Petrobras (Brazil), each made
Group. PDVSA (Venezuela), Petrobras (Brazil), and Pemex more significant discoveries in the 1980s in their
(Mexico) are responsible for most of the major discoveries countries than the combined discoveries of all the other
during this period. multinational companies in all of Latin America. This
most likely reflects the quality of reserve areas to which
the NOCs maintain exclusive exploration rights.
Oil and Gas Discoveries and Basin Resource Predictions in Latin America 57

Table 1—Largest Oil Fields in Latin American Countries, Table 2—Largest Gas Fields in Latin American Countries,
1980–1992 1980–1992

BBO tcf
Country Field Year Operator (est.) Country Field Year Operator (est.)
Brazil Marlim 1985 Petrobras 5.40 Peru Cashiriari 1986 Shell 10.00
Mexico Uech 1986 Pemex 2.00 Brazil Marlim 1985 Petrobras 7.70
Venezuela Furrial- Colombia Cusiana 1988 BP 6.50
Musipan 1986 PDVSA 1.90 Venezuela Furrial-
Colombia Cusiana 1988 BP 1.70 Musipan 1986 PDVSA 5.65
Peru Cashiriari 1986 Shell 0.50 Argentina Ara-Hidra 1981 Total 0.98
Ecuador Ishpingo 1992 PetroEcuador 0.23 Mexico Cardinas 1980 Pemex 1.00
Argentina El Trapiel 1991 San Jorge 0.17 Trinidad Chaconia 1981 Tenneco 0.75
Trinidad NE Soldado 1981 Texaco 0.10 Bolivia Surubi 1992 Maxus 0.25
Surinam Tambarejo 1980 Statsole 0.06 Chile Lago
Mercedes 1992 Texaco 0.05

Based on the largest resources discovered per wildcat


well, the most successful multinational companies have
been Occidental, Royal Dutch/Shell, British Petroleum,
Total, and Conoco (Figure 6). A second group of
companies, including Mobil, ARCO, Chevron, Amoco,
Exxon, Elf, and Texaco, each found less than 5 MMBOE
per exploration well drilled.
Interestingly, the two most successful companies,
Occidental and Royal Dutch/Shell, have each partici-
pated in one major discovery (>500 MMBOE), Caño
Limón and Cashiriari, respectively, during the last
decade. Even without the benefit of this major discovery,
Occidental would still be among the more successful
companies. However, Royal Dutch/Shell would fall into
the second tier of companies.

Figure 6—Number of wildcat wells drilled in Latin America


BASIN RESOURCE ESTIMATES versus net oil found (by company) for 12 MNCs,
1980–1992. The points plotted on this graph give a semi-
The purpose of this analysis is to assess the prospec- quantitative assessment of the technical and oil-finding
tivity of producing basins in order to establish priorities success rate in each of the countries studied. Values are
for exploration. In 1989, Amoco began interpreting the plotted twice for Royal Dutch/Shell and Oxy to show the
exploration maturity of a basin based on the field size effect that Cashiriari and Caño Limón had, respectively, on
distribution and the exploration history of discoveries the overall oil-finding success rate of each company.
using a methodology similar to that of the U.S. Geolog- (Wells drilled in Mexico, Cuba, and Brazil are not included.)
ical Survey (Dolton et al., 1981; Mast et al., 1989; Drew
and Shuenemeyer, 1993). Ideally, a minimum of 40
discoveries are required to do a rigorous trend analysis this evaluation were established by Amoco’s Exploration
(5% of 20 is one field). For this reason, we have applied Basin Analysis Group. Resource information from
the statistical techniques toward emerging or mature commercial databases, as well as other scout informa-
producing basins rather than frontier areas. tion, was used to analyze the discoveries.

Historical Field Size Analysis Exploration Efficiency


Historical field size analyses are estimates of the Exploration efficiency is a measure of the historical
remaining undiscovered field sizes in a basin. This statis- resource discovery success in a given basin. Tradition-
tical analysis may be useful in focusing exploration ally, the largest fields are discovered early in an effi-
opportunities. The criteria for estimating a probability ciently explored basin, and additional discoveries are
that undiscovered fields exist is based on a combination generally smaller through time. A basin with a 100%
of basin analysis plots. These include exploration effi- exploration efficiency is demonstrated graphically on
ciency, the 95th percentile of the most recent discoveries curve A in both graphs in Figure 7. In an inefficiently
from field size population statistics, and field size distrib- explored basin, large and small fields are discovered
ution plots. The population statistical techniques used in randomly. Curve C in Figure 7 shows an exploration effi-
58 Kronman et al.

(a)

Figure 8—Comparison of (a) worldwide (excluding the


United States and Canada) and (b) Latin American field
sizes (median and 95th percentile) grouped by “sets” of
chronologically ordered discoveries. In (a) , set 1 contains
the first 400 discoveries, followed by set 2 which contains
the second 400 discoveries. In (b), Latin American fields
(b) have 200 discoveries per set. A comparison of these two
graphs shows that the median and 95th percentile field size
declines over time.
Figure 7—Exploration efficiency graphs for the (a)
Maracaibo basin and (b) Maturin basin, Venezuela. The
Maracaibo is an example of a basin with a high exploration Latin American basins can be divided into a two-
efficiency (54.3%), while the Maturin has a low exploration tiered classification of high (>50%) and low (<50%)
efficiency (15.5%). Discoveries are listed chronologically exploration efficiencies. The Middle Magdalena, San
along the horizontal axis, and their corresponding field Jorge, Marañon-Napo-Putumayo, Maracaibo, and
sizes (vertical axis) are plotted cumulatively (curve B). An Reforma-Campeche basins are in the high tier, sug-
idealized cumulative field size is shown in each graph gesting that the bigger fields were found early in the
(curve A), which assumes that the largest fields are discov- exploration of these basins. The Neuquén, Tampico-
ered first in a basin and that smaller additional discoveries
Misantla, Llanos, South Basin (Trinidad), Austral, Upper
are made over time. Curve C represents a hypothetical
basin with a 0% exploration efficiency. Magdalena, Maturin, and Campos basins have the
lowest exploration efficiencies, suggesting that large
fields remain to be discovered there.
ciency of 0%. The actual exploration efficiency value is a
measure that compares a basin’s exploration history Field Size Population Statistics
(curve B, Figure 7) with the idealized efficiently explored Field size population graphs record the change in field
basin (curve A). A high exploration efficiency implies size discoveries through time. Known fields are consecu-
that the largest fields in a basin have already been found tively ordered and chronologically grouped in sets and
and that a new play type or introduction of new tech- plotted against the log of the field sizes discovered.
nology is needed to find additional large fields, such as Various size percentiles for discoveries are calculated for
the Maracaibo basin (Figure 7a). A low exploration effi- each set. The 95th percentile trend is significant because
ciency, such as the Maturin basin (Figure 7b), indicates it suggests the top 5% of the largest field sizes yet to be
the possibility of discovering large fields because existing discovered.
play types may not have been fully exploited. Low The worldwide field size statistics for liquid discov-
exploration efficiencies may be the result of successful eries, excluding the United States and Canada, illustrates
new technology or a play concept introduced late in the the overall decrease in the 95th percentile or the top 5%
history of the basin, licensing policy, or physical accessi- of the discoveries through time (Figure 8a). These discov-
bility such as drill depth or geopressure. eries are grouped in sets by consecutive discoveries of
Oil and Gas Discoveries and Basin Resource Predictions in Latin America 59

(a)

Figure 9—The 95th percentile of field sizes for selected


Latin American basins, in MMBO. The 95th percentile is
extrapolated to predict the top 5% of the next set of discov-
eries for each respective basin.

400 fields. The 95th percentile decreases sharply from 1


BBO prior to about 1970, to 650 MMBO in the 1970s, and
to 150 MMBO after 1980. Figure 8a shows that it is
increasingly more difficult to discover giant fields greater
than 500 MMBO. The Latin American discovery size (b)
statistics, grouped in sets of 200 discoveries, show a
similar trend in the decrease in the 95th percentile over
time (Figure 8b). The 95th percentile also starts at 1 BBO,
falls to about 600 MMBO in the 1970s, and decreases
further to 100–250 MMBO in the 1980s.
The most important basins are ranked by their 95th
percentile field size in Figure 9. These 95th percentiles can
be extrapolated to predict the top 5% of the next set of
discoveries. Campos, Maturin-Furrial, Llanos, Maracaibo,
and Reforma-Campeche have 95th percentiles greater
than 1 BBO. Trinidad has 600 MMBO, and Tampico-
Misantla, Marañon-Napo-Putumayo, and Neuquén have
95th percentiles between 100 and 250 MMBO.

Field Size Distribution (c)

Arps and Roberts (1958) were the first to report a


lognormal distribution of oil field sizes with respect to
frequency above an economic threshold on the eastern
flank of the Denver basin. Other mature basins have also Figure 10—Theoretical population distribution of discov-
shown a lognormal distribution of field sizes with eries by field size class in a basin. (a) Field size distribu-
respect to frequency, such as the Western Canada sedi- tions in basins generally approach a lognormal pattern.
mentary basin (McCrossan, 1969; Kaufman et al., 1975), Smaller fields commonly fall below the economic “trunca-
the Gulf of Mexico (Schuenemeyer and Drew, 1983), and tion” or threshold. An exponential decrease is generally
the Permian basin (Schuenemeyer and Drew, 1983). seen in the number of fields as field size increases. An
Undiscovered field sizes in a basin can be predicted underrepresented field size class along the lognormal
curve suggests that additional fields of that field size are
on the basis of theoretical field size population distribu- yet to be found. (b) Field size distribution for the Maturin
tion and the exploration history of discoveries within the basin (Venezuela) prior to 1980. Based on this graph, the
basin. The overall distribution of hydrocarbon fields is number of fields in those field size classes with an arrow
lognormal for a large number of smaller fields and a over them are believed to be underrepresented. (c) Addi-
decreasing number of larger fields (Figure 10a). The tional field discoveries made in the Maturin basin after
economic truncation or avoidance toward finding 1980, many of which occurred in the underrepresented
smaller fields causes the observed population to be field size classes.
60 Kronman et al.

Basins with estimated undiscovered recoverable


resources exceeding 5 BBO are the Maturin, Reforma-
Campeche, Maracaibo, and Campos basins. A large
second tier of basins includes the Llanos, Ucayali,
Magdalena, Marañon-Napo-Putumayo, and Neuquén
basins, with undiscovered recoverable resources ranging
from about 1 to 5 BBO. A third tier of basins includes the
Tampico-Mislanta, Austral, Isthmus-Saline, San Jorge,
and South Basin (Trinidad), with remaining undiscov-
ered recoverable resources ranging from 0.5 to 1 BBO.
There are estimated to be 40–80 BBO of remaining recov-
erable resources in this group of 15 Latin American
basins. Approximately 170 BBO of ultimate reserves
(cumulative and proven) are estimated to occur in the
same basins.
The values for these undiscovered field size estimates
and resources must be used with due caution. There is an
obvious danger of overestimating a basin’s potential and
an even greater danger of ignoring a basin where the
next giant discovery has yet to be made. Mature basins
whose resource distribution is allegedly well understood
may offer new opportunities when a new play concept,
technological breakthrough, or political change opens
important new trends. In contrast, frontier basins
generally have too few data for rigorous analysis.
Figure 11—Map of Latin America showing estimated
remaining discovery sizes by class for the major petroleum
basins. The estimated remaining discovery sizes are based
largely on field size statistical analysis. Frontier basins, or
CONCLUSIONS
those basins with limited data, are not included.
From 1980 to 1990, Latin America had an estimated
17% success rate in the more than 4500 wildcats drilled.
This percentage, as well as the percentage of significant
skewed in favor of larger fields. Some techniques use oil and gas discoveries (>100 MMBO) and giant oil and
these lognormal distributions to estimate undiscovered gas fields (>500 MMBO), is comparable to the success
resources and potential missing field sizes (Schuene- rates for the 19,790 wildcats drilled worldwide over the
meyer and Drew, 1983). For example, Figure 10b same time period (excluding the United States and
displays the field size distribution in the Maturin basin Canada).
prior to 1980. Arrows denote possible missing field sizes In the 1980s, ten major discoveries containing more
that fall below the lognormal curve. The fields discov- than 1 BBO each were made in four countries: Brazil
ered after 1980 fill in the predicted distribution pattern (three), Mexico (three), Venezuela (two), and Colombia
(Figure 10c). Undiscovered field sizes reinforce and (two). These four countries account for 87% of the 70
increase the confidence for prospects of a predicted field significant discoveries greater than 100 MMBO. Three
size. NOCs—PDVSA of Venezuela, Petrobras of Brazil, and
Basins that have a high probability of containing addi- Pemex of Mexico—have made 35 discoveries, four times
tional fields greater than 500 MMBO (Figure 11) are the the number of discoveries made by the MNCs combined.
Maracaibo, Campos, Maturin, Llanos, and Reforma- This most likely reflects the quality of the basins to which
Campeche. Smaller but potentially economic discoveries the NOCs have exclusive rights.
in the 50–250 MMBO range may yet exist in the Historical field size analysis based on exploration effi-
Tampico-Misantla, Neuquén, San Jorge, Astral, Tarija, ciency, 95th percentiles, and historical field size distribu-
Marañon-Napo-Putumayo, and Magdalena basins. tion indicates that significant economically attractive
opportunities may still exist in Latin America. The 95th
Prediction of Undiscovered Resources percentile, or the top 5%, of the most recent discoveries in
all of Latin America is 100–250 MMBO.
Estimates of the number and size of fields remaining Potential giant discoveries greater than 500 MMBO
to be discovered that are above an economic threshold may yet exist in the following basins:
are the undiscovered resources for that basin. These
estimates are integrated with the exploration efficiency • Campos basin in Brazil
and field size population statistics before being evaluated • Llanos basin in Colombia
by experts familiar with the drilling history and hydro- • Maturin and Maracaibo basins in Venezuela
carbon system of the basin. • Reforma-Campeche basin in Mexico
Oil and Gas Discoveries and Basin Resource Predictions in Latin America 61

Smaller but potentially economic discoveries in the recoverable conventional resources of oil and gas in the
50–250 MMBO range possibly remain to be made in the United States: USGS Circular 860, 87 p.
following basins: Drew,L. J., and Schuenemeyer, J. H., 1993, The evolution and
use of discovery process models at the United States
Geological Survey: AAPG Bulletin; v. 77, p. 467-478.
• Neuquén, San Jorge, and Austral basins in
Kaufman, G. M., Y. Balcer, and D. Kruyt, 1975, A probabilistic
Argentina model of oil and gas discovery, in J. D. Haun, ed., Methods
• Tarija basin in Bolivia and northern Argentina of estimating the volume of undiscovered oil and gas
• Marañon-Napo-Putumayo basin in Ecuador, resources: AAPG Studies in Geology 1, p. 113–142.
Peru, and Colombia Kronman, G. E., A. M. Aleman, S. W. Rushworth, 1993, Oil
• Tampico-Misantla basin in Mexico discoveries and basin resource predictions in Latin
• Magdalena basin in Colombia America: past, present, and future: AAPG Bulletin, v. 77,
p. 329.
Based on the statistical techniques used in this study, Kronman, G. E. and K. D. Smith, 1993, The rise of National Oil
about 40–80 BBO of undiscovered recoverable resources Companies: the new Seven Sisters?: 78th Annual AAPG-
SEPM Convention, April, New Orleans, Louisiana,
remain in the basins analyzed. These basins at present
Abstracts with Program, p. 132.
have approximately 170 BBO of proven and probable Mast, R. F., G. L. Dolton, R. A. Crovelli, D. H. Root, E. D.
reserves. Attanasi, P. E Martin, L. W. Cooke, G. B. Carpenter, W. C.
Pecora, and M. B. Rose, 1989, Estimates of undiscovered
conventional resources of oil and gas in the United
States—a part of the nation’s energy endowment: USGS
Acknowledgments Many colleagues at Amoco have and Minerals Management Service, 44 p.
McCrossan, R. G., 1969, An analysis of size frequency distrib-
contributed to this study: Josh Rosenfeld, Carol Kazmer, Bob
ution of oil and gas reserves of western Canada: Canadian
Erlich, Don Felio, Larry Parks, Richard Steinmetz, Joe Journal of Earth Sciences; v. 6, p. 201–211.
Malagowicz, Peter Carragher, Lee Distefano, Don Regan, Jon Schuenemeyer, J. H., and L. J. Drew, 1983, A procedure to
Blickwede, Robert Marksteiner, Eric Green, Bob Harper, Mike estimate the parent population of the size of oil and gas
Deming, Claudio Manzolillo, Wendy Hale-Erlich, Marcos fields as revealed by a study of economic truncation: Math-
Roberto, Wolfgang Schollnberger, and Grant Emms. Informa- ematical Geology; v. 15, p. 145–161.
tion was also provided by Amoco’s Resource Assessment and
Exploration Analysis groups. We thank Anthony J. Tankard,
Peter Aukes, and Robert A. Meneley for their excellent review
of this paper. Technical help from Lori Fortner, Joe Tunnell,
John Cain, Zoila Torres, and Jeff Suiter and drafting by Teresa
Holmes is also appreciated. We would also like to thank Amoco
for permission to publish this paper.

Authors’ Mailing Address


REFERENCES CITED George E. Kronman
Sandra W. Rushworth
Arps, J. J., and T. G. Roberts, 1958, Economics of drilling for Keith Jagiello
Cretaceous oil on the east flank of Denver-Julesburg basin: Antenor Aleman
AAPG Bulletin, v. 42, p. 2549–2566. Amoco Production Company
Dolton, G. L., K. H. Carlson, R. R. Charpentier, A. B. Coury, R. 501 WestLake Park Boulevard
A. Crovelli, S. E. Frezon A. S. Khan, J. H. Lister, R. H. Houston, Texas 77253-3092
McMullin, R. S. Pike et al., 1981, Estimates of undiscovered U.S.A.
Petroleum Basins of Southern South America:
An Overview

C. M. Urien J. J. Zambrano
Buenos Aires Technological Institute Regional Ground Water Institute
Buenos Aires, Argentina San Juan, Argentina

M. R. Yrigoyen
National Academy of Sciences
Buenos Aires, Argentina

Abstract

F rom the Cambrian to Late Jurassic, the basins and arches of southern South America were oriented
approximately north-south. Subsequently, southwest-northeast trending stresses related to the breakup
of Gondwana and the opening of the South Atlantic imposed new structural alignments, the effects of which
were widespread. These two tectonic regimes encompass six stages of basin formation:

• Cambrian–Middle Devonian terrigenous clastics, carbonates, and intrusives occurred along the
western edge of the Brazilian, Puna, and Pampas shield areas.
• Carboniferous–Late Jurassic sedimentation in the intracratonic rifts was mainly of continental origin.
Marine clastics accumulated in a foreland basin in front of a volcanic arc on the western edge of the
continent. This period ended with the Late Jurassic breakup of Gondwana and extrusive magmatism.
• Late Jurassic extension marked by widespread marine flooding affected vast regions of the continent.
Clastics, evaporites, and acidic volcanics constitute the “Andean foreland” succession which was asso-
ciated with a volcanic arc system. In Patagonia, acidic volcanics covered the North Patagonia massif
and Deseado craton and earlier basins.
• After breakup, sedimentary prisms formed along the Atlantic margin on the western margin. Thick
sequences of shales, limestones, evaporites, and pyroclastics were associated with middle–Late Creta-
ceous volcanic arc and back-arc settings. New acidic intrusions and the Andean batholith are also
dated to this period.
• The Late Cretaceous–early Tertiary (Laramide) was marked by development of the Andean fold and
thrust belt and final emplacement of the Andean batholith. A flexural foreland basin formed in front
of the Andes. Passive margin sedimentation dominated the eastern margin.
• The Tertiary was a time of Andean mountain building and passive margin subsidence. The thrust belt
supplied thick sedimentary fills to the foreland basin. Shallow transgressions covered much of
Patagonia and the Pampa plains.

This tectonic evolution is expressed in a complex array of composite basins. These tectonic, structural, and
depositional patterns were also responsible for a suite of petroleum systems, many of them commercially
significant.

Resumen

D esde el Cámbrico hasta el Jurásico, las cuencas y arcos estructurales se orientaban aproximadamente
Norte-Sur. Posteriormente, como resultado de la fractura del Gondwana y la apertura del Atlántico
Sur, se produjeron esfuerzos con tendencia predominante Sudoeste-Noreste, que impusieron un nuevo orde-
namiento estructural cuyos efectos fueron de gran extensión. Estos dos regímenes tectónicos abarcan seis
etapas en la formación de las cuencas:

• Desde el Cámbrico hasta el Devónico Medio, se distribuyen clásticos terrígenos, carbonatos e intru-
sivos a lo largo del borde occidental de los escudos de Brasil, Puna y Pampeano.
• La sedimentación en el Jurásico Tardío en los rifts intracratónicos, fue principalmente de origen conti-
nental. En una cuenca antepaís, cerrada por el arco volcánico en el borde occidental del continente, se
acumularon sedimentos clásticos marinos. Este período concluyó con la fractura del Gondwana y el
magmatismo extrusivo del Jurásico Tardío.

63
Urien, C. M., J. J. Zambrano, and M. R. Yrigoyen, 1995, Petroleum basins of southern
South America: an overview: in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum
basins of South America: AAPG Memoir 62, p. 63–77.
64 Urien et al.

• Extensión tectónica tensional durante el Jurásico Tardío, señalada por extensas invasiones marinas que
inundan vastas regiones del continente. Clásticas, evaporitas y vulcánitas ácidas constituyen la
sucesión de la cuenca andina de antepaís, relacionada con un sistema de arco volcánico. En la
Patagonia, vulcánitas ácidas cubrieron el Macizo Norpatagónico, el cratón Deseado y cuencas prece-
dentes.
• Luego de la ruptura continental, se formaron prismas sedimentarios a lo largo del margen occidental y
gruesas secuencias de esquistos, calizas, evaporitas y piroclásticas se asociaron a escenarios de arcos y
retro-arcos volcánicos del Cretácico Medio y Tardío. Nuevas intrusiones ácidas y el emplazamiento del
batolito Andino también datan de este período.
• El Cretácico Tardío/Terciario Temprano (Larámico) se caracterizó por el desarrollo de la Faja Andina
de Plegamientos-Corrimientos y el emplazamiento final del batolito Andino. Una cuenca flexural
antepaís se forma frente a los Andes. En el borde oriental predominaba una sedimentación de margen
pasivo.
• El Terciario fue época de elevación de las montañas de los Andes y de subsidencia del margen
atlántico pasivo. La Faja de corrimiento aportó gruesos rellenos sedimentarios a la cuenca del antepaís.
Transgresiones someras cubrieron gran parte de la Patagonia y las llanuras Pampeanas.

Esta evolución tectónica tiene expresión en un complejo sistema de cuencas compuestas. Esta configuracion
tectónica, estructural y sedimentaria fue también responsable de una serie de sistemas petrolíferos, muchos de
ellos comercialmente significativos.

INTRODUCTION GEOLOGIC SETTING


The southern part of the South American continent In southern South America, the upper Paleozoic,
has an arrangement of stable cratonic blocks and sedi- Mesozoic, and Cenozoic sedimentary cover overlies older
mentary basins, which suggests that deformation and basement, to a large extent early–middle Paleozoic in age.
basin formation have reactivated old terrane boundaries. In central western Argentina and especially Patagonia
However, this tectonic–structural framework has been (Figure 1), the basement has a variety of stratigraphic and
largely obscured by subsequent magmatic events. structural characteristics, but it is generally too young to
Nevertheless, the broadscale architecture of the conti- have been cratonized. It comprises mainly metasedimen-
nental platform is understood. Since the Precambrian, tary and igneous assemblages.
this platform has been modified by tectonism and sedi- The age of the intrusive and extrusive igneous rocks
mentation on the margin of the old cratonic core. This varies from Paleozoic to Late Jurassic. These igneous
basement and its lower Paleozoic cover built the foreland complexes form large massifs (Figure 1). The North
complex onto which successive fore-arc and volcanic arc Patagonia massif (or Meseta de Somuncurá of some
tracts and the Andean orogenic belt have encroached authors) is a complex of intrusives and extrusives that
(Figure 1). resulted from successive magmatic episodes, dating from
The southern part of the South American continent Carboniferous to Permian–Triassic (Stipanicic, 1969a;
underwent a complex evolution, reflecting interaction of Pankhust et al., 1992; Rapela and Pankhurst, 1992).
this basement framework with several episodes of Farther south is the Deseado massif, which forms the
Cambrian–Jurassic tectonism, followed by widespread northwestern sector of the larger Deseado-Malvinas
extension related to continental fragmentation and platform, a suite of intrusive uplifted blocks that are
opening of the Atlantic (Uliana and Biddle, 1988). Many covered by Mesozoic and Cenozoic volcanics (De Giusto
of those basins are composite in form and have several et al., 1980). They combined with other basement blocks
hydrocarbon habitats. to build a series of fault-bounded ridges, between which
The most prolific hydrocarbon-bearing basins are several sedimentary depocenters occur beneath the
those forming the sub-Andean belt. These are a suite of Jurassic volcanic cover.
long-lived composite basins that have been subjected to a Therefore, south of the west Pampean region (south of
foreland style of subsidence only since the Late Creta- 38˚ S lat.), the “basement” generally ranges in age from
ceous. Most of the hydrocarbons are produced from late Paleozoic to early–middle Mesozoic (Forsythe, 1982;
Mesozoic sequences (Upper Triassic, Jurassic, and Creta- Ramos et al., 1982; Gust et al., 1985). The topography of
ceous), with local production from the Paleozoic section this continental platform includes several conspicuous
(Silurian–Devonian and Carboniferous) and subordinate elements (Figure 1):
Tertiary reservoirs.
The aim of this paper is to review the evolution of • The Pampean-Transpampean-Puna arch formed
these hydrocarbon-bearing basins through several on Precambrian and Cambrian basement (Bracac-
distinct tectonic episodes. We will discuss the petroleum cini, 1960; Amos, 1972; Aceñolaza et al., 1982).
systems on this basis and will speculate on some • To the west, an early Paleozoic magmatic arc is
untested opportunities. The intention is to present a preserved in the frontal Cordillera and the San
“thumbnail sketch” of this complex history. Rafael uplift (Criado Roque and Ibañez, 1979).
Petroleum Basins of Southern South America: An Overview 65

Figure 1—Architecture of southern South America, including basins, arches, and massifs.
66 Urien et al.

• The North Patagonia and Deseado massifs were


formed more recently by intense magmatic A suite of paleogeographic reconstructions spanning
activity. Although magmatism began in the late Paleozoic–early Tertiary time are shown in Figures 3
Carboniferous, the principal activity occurred in to 8. These maps include the intrusive and extrusive
the Late Jurassic. igneous rocks.
• An acidic volcanic arc formed along the Andean The Choiyoi basement of the Cuyo basin (Figure 1)
deformation belt. However, it did not stop marine consists of Upper Permian–Triassic effusives and pyro-
access to the back-arc basins of the region now clastic rocks (Rolleri and Fernandez Garrasino, 1979).
caught up in the main Cordillera. Cambrian–Lower Permian deposits are also known. In
the Bolsones basins north of Cuyo, upper Paleozoic–
The western and southern parts of the region were Triassic stratigraphy is approximately continuous
constructed by accretion of successive volcanic and sedi- without acid volcaniclastic rocks intervening; this succes-
mentary sequences along the margin of a dissected sion overlies Precambrian basement.
cratonic platform. This succession was deformed in The Neuquén contains a Jurassic–Cretaceous succes-
several phases of Andean orogeny, mainly in the middle sion above an economic basement of Permian–Triassic
and late Tertiary. Throughout the Paleozoic and Choiyoi Group volcanics. Paleozoic and Triassic
Mesozoic, these basins developed as intracontinental rifts sediments are believed to occur in the northeastern part
or sags and were separated from the Pacific margin fore- of the basin, suggesting that the basement may even
arc system by marginal platforms (see Lister et al., 1986). contain Devonian or Paleozoic intrusives and metamor-
They were not totally open to the Pacific. Marine and phic rocks.
continental sedimentary successions alternate (Charrier In the San Jorge basin of central Patagonia (Figure 1),
and Vicente, 1972; Auboin et al., 1973). Most of the the Upper Jurassic–Cretaceous succession overlies
marine incursions were of a restricted nature (euxinic), as Middle–Upper Jurassic volcanic and pyroclastic rocks of
reflected in the accumulation and preservation of the Lonco Trapial Group and Chon Aike Formation
petroleum-prone source rocks. In contrast, the (Lesta et al., 1980). A complex stratigraphy of Carbonif-
nonmarine terrigenous clastics formed profilic reservoirs erous, Permian, and Liassic rocks has been identified
in a succession of stacked sedimentary wedges. beneath this volcanic assemblage (Piatnitzky, 1933;
The composite nature of these basins, their long and Suero, 1948; Ugarte, 1966). Basement is an early Paleozoic
complex tectonic history, the periodic marine incursions, accretionary terrane. In the central parts of the San Jorge
and the stacked reservoir intervals suggest a variety of and Magallanes (Austral) basins, the Middle–Upper
play types and exploration opportunities spanning the Jurassic volcanics and pyroclastics of the El Quemado
entire stratigraphic column. It is these petroleum and Tobífera formations cover a series of north-south
systems, tested and untested, that are the focus of this trending half-grabens of Jurassic age that contain a conti-
paper. nental fill.
There are four principal families of basin-forming The Magallanes (Austral) and Malvinas basins
basement: (Figure 1) contain an Upper Jurassic and Lower Creta-
ceous sedimentary fill above Middle Jurassic acidic
1. First are Precambrian metamorphic rocks and volcaniclastic rocks and rift-like Lower Jurassic conti-
intrusives and Cambrian–Ordovician metasedi- nental deposits. The upper Paleozoic basement consists
mentary and intrusive rocks (Baldis and of pyroclastic and metamorphic lithologies (Riccardi and
Bardonaro, 1984; Prozzi and Rosso, 1990). Taconic Rolleri, 1980).
tectonism closed this period (Figure 2a).
2. Silurian–Devonian clastic sedimentary cover was
subjected to low-grade metamorphism (Figure 2b). PHANEROZOIC BASIN EVOLUTION
These sequences were influenced by the Ocloyic
and Chanic diastrophisms (broadly equivalent to The sedimentary basins of southern South America
the Caledonian) (Cuerda and Furque, 1979). preserve a complex record of Phanerozoic evolution,
3. Volcanic rocks and plutons form a prominent repeatedly reactivating old lineaments in a variety of
Permian–Triassic magmatic arc associated with regional stress settings. We will discuss this evolution in
continental and marine sediments. These were later three principal phases. First, the Paleozoic–Jurassic
deformed by Carboniferous–Early Permian evolution established the cover sequence of south-
(Variscan) tectonism (Figure 2c) (Groeber et al., western Gondwana. Second, major diastrophism in the
1952; Caminos, 1979a; Yrigoyen, 1979; Gust et al., Late Jurassic (Kimmeridgian, late Malm) produced an
1985; Pankhurst et al., 1992). unconformity in most of these basins. This is locally
4. The Late Triassic to Middle–Late Jurassic migration known as the Auracanian orogeny, which initiated a new
of the magmatic arc toward the west is recorded in cycle of basin behavior and established the present distri-
the pluton emplacements, volcanism, and volcani- bution of basins. Third, latest Jurassic, Cretaceous, and
clastic rocks that are interbedded with continental Cenozoic evolution included the extensional regime that
and shallow marine sequences (Figure 2d) culminated in the Atlantic opening, as well as the
(Stipanicic and Rodrigo, 1969; Stipanicic and Laramide and Andean mountain building episodes. This
Linares, 1975; Franchi et al., 1989). final phase established the modern subcontinent.
Petroleum Basins of Southern South America: An Overview 67

Figure 2—Paleogeography of southern South America through four principal basin-forming stages: (a) Cambrian–Ordovician,
(b) Silurian–Devonian, (c) Carboniferous–Early Permian, and (d) Late Permian–Triassic. BS, Brazilian shield; NPM, North
Patagonia massif; PR, Pampean Ranges; PRC, Plata-Ribeira craton; TPR, Trans-Pampean Ranges; DM, Deseado massif.
68 Urien et al.

PALEOZOIC–MIDDLE MESOZOIC are preserved. The conjugate Cape basin of South Africa
was part of this Ordovician–Early Devonian depositional
DYNAMICS landscape (Tankard et al., 1982).
Early Paleozoic Basins Petroleum Geology
Pre–Middle Ordovician The Chaco and Tarija basins contain Silurian and
Devonian source rocks. The Silurian sequence is known
The western selvage of the subcontinent south of 25˚ S as the Lipeón and Copo formations. However, most oil
lat, spanning Bolivia, Argentina, and Chile, was and gas generation is attributed to the Devonian Los
generated at the beginning of the Paleozoic by accretion Monos shales, although the Santa Rosa and Tonono
of a magmatic arc and sedimentary wedges onto a formations also have potential. The estimated recover-
Precambrian core. The basement core is relatively young able reserves are 1.1 billion bbl of oil and 18.4 tcf of gas,
(1800–570 Ma) and forms the Brasiliano platform (Brito the Argentinian share of which is about 380 million bbl
Neves and Cordani, 1991), or central cratogene of Brac- of light oil and 12 tcf of gas (Yrigoyen, 1991a). In the
cacini (1960). This tectonized terrace wedge formed west Altiplano and the northwestern Paraná basin, there are
of the Pampean cratonic block and Puna arch and was coeval deposits that may have petroleum potential.
bordered westward by oceanic crust. Silurian–Devonian flysch-like sediments are also known
This early Paleozoic passive margin succession was in the Precordillera of San Jorge and Mendoza and the
built by deep water turbidite sedimentation over which a San Rafael uplift. Their petroleum-generating potential is
clastic platform prograded (González Bonorino and unknown.
González Bonorino, 1991). Distally, this prism is believed
to have encroached on oceanic crust. Remnants of this
basin are preserved in the Altiplano and Eastern
Late Paleozoic Basins
Cordillera of Bolivia and northwestern Argentina and in Devonian–Early Carboniferous
the Pampean Ranges. The grade of metamorphism
increases southward. During Late Devonian–Early Carboniferous time,
intense deformation folded, and in places metamor-
Middle–Late Ordovician phosed, the lower Paleozoic succession. This episode is
colloquially known as the Chanic orogeny and is
In the Middle Ordovician, igneous intrusive bodies expressed regionally as a major unconformity. Early
were emplaced in the cratonic crust, the erosion of which Carboniferous subsidence was generally vigorous, but
is believed to have supplied voluminous clastics to the reactivated old trends (Figure 2c).
Ordovician basins (Furque and Cuerda, 1979). The The earliest deposits occur in the Tarija basin of
Buenos Aires hills contain contemporaneous deposits of southern Bolivia and northwestern Argentina where they
a fringing shallow marine platform (Harrington, 1956; are called the Machareti Group. Equivalent deposits in
Andreis et al., 1989). the Chaco-Paraná basin form the Sachayoj Formation
(Padula and Mingramm, 1969; Pezzi and Mozetic, 1989).
Late Ordovician–Middle Devonian Elsewhere, broadly contemporaneous deposits occur in
The Late Ordovician–Middle Devonian episode began isolated depocenters. Examples include the El Toco
with the earliest evidence of collision tectonics when a Formation of northern Chile, the Malimán Formation in
magmatic arc was accreted to the continental margin. Its the Argentinian Precordillera (Scalabrini Ortíz, 1972), and
remnants occur in the western Pampean Ranges. less well understood deposits of central and southern
Intrusive activity dominated the Famatina mountain Patagonia, such as the Tepuel Group and the Bahia de la
ranges. In the northern Chilean coastal Cordillera and the Lancha Formation (Riccardi and Rolleri, 1980).
eastern Pampean Ranges, plutonism was less extensive In the middle–Early Carboniferous (Visean), right-
(Gordillo and Lencinas, 1979; Pichowiak et al., 1987). The lateral wrenching dissected the Pampean basement into
far-flung effects of this plate margin collision formed a a suite of pull-apart basins (Figure 3) (Fernandez-Seveso,
remarkable intracontinental or epeiric basin in the Early 1995). These basins were initially isolated. Deposition
Silurian. This shallow basin complex once covered occurred in coarse alluvial systems that were stacked
Bolivia, northern Argentina, and Paraguay, where it is against active boundary faults and in finer grained lacus-
known as the Chaco, Tarija, and Chaco-Paraná basins, trine and shallow marine settings. A periglacial climate is
respectively. Anoxic environments attributed to suggested by the varved shales and dropstones. Eyles
restricted circulation formed organic-rich source rocks, (1993) suggests that glacial conditions were a response to
especially in the Devonian Los Monos Formation (Pareja tectonism and uplift. The pull-apart basins in central
et al., 1978; Pezzi and Mozetic, 1989). western Argentina are known as bolsones. By the end of
This sequence forms a sedimentary belt in the Puna the Carboniferous and Permian, these fault-controlled
and Pampean ranges, extending from the Altiplano to basins were joined together in a broad region of subsi-
northwestern Patagonia (Figure 2b). It also covers the dence (see Fernandez-Seveso, 1995, for a detailed
continental interior as far as Malvinas Island (Turner, description). In Patagonia, sedimentary rocks of this age
1972; Pareja et al., 1978). In the Southern Hills of Buenos rest on early Paleozoic basement (Uliana and Biddle,
Aires, Silurian–Lower Devonian shelf clastic sequences 1987, 1988). These Carboniferous–Permian basins were
Petroleum Basins of Southern South America: An Overview 69

Triassic basins were cut off from marine circulation.


Permian–Triassic volcanics and volcanoclastics beneath
these intracontinental rifts are assigned to the Choiyoi
Group.

Petroleum Geology
Potential Carboniferous–Permian source rocks have
been reported from many basins, but so far appear to
have attracted little commercial interest. In the Tarija
basin of northwestern Argentina and Bolivia, the Upper
Permian Vitiacuá Formation is an oil-prone lacustrine
mudstone. Its counterpart in the Paraná basin of Brazil is
the Iratí Formation, which locally averages about 10%
total organic carbons (TOC) and has been mined.
In the intermontane basins of central western
Argentina, lagoonal and lacustrine organic-rich carbona-
ceous shales (e.g., Patquía Formation of the Paganzo
basin) contain type I and II kerogens. Numerous oil
seeps occur along old strike-slip fault zones. The San
Rafael uplift may contain similar petroleum potential.
Source rocks and reservoirs rocks are known. The risk
would appear to be maturation and seals.
Thick upper Paleozoic shallow marine and deltaic
sedimentary rocks are exposed in the Pampa de Agnia
basin of southern Patagonia. These rocks are believed to
extend southward to the western San Jorge basin. We
speculate that the offshore Colorado basin may have
similar potential considering its proximity to Upper
Permian Iratí equivalents in Uruguay.

Triassic Rift Basins


Basin Development
The Late Permian–Early Triassic landscape of south-
western Gondwana was disrupted by an orogenic
episode that deformed and inverted many of the late
Paleozoic basins. This deformation is expressed in the
Cape foldbelt of South Africa and in the Sierras de la
Ventana and Pintada of Argentina. This is not a contin-
uous mountain belt, but has several offsets that reflect
accommodation zones between the original basins. An
important consequence of this deformation was the
Figure 3—Paleogeography of southern South America postorogenic relaxation of compressive stresses and the
during the late Paleozoic. The legend shows the distribu- formation of a suite of extensional basins by orogenic
tion of depositional environments. collapse (Figure 4). Extension persisted intermittently
into the Early Jurassic (Stipanicic, 1969b). These Triassic
inverted by Tertiary deformation of the Andean era, basins were essentially landlocked and contain thick
forming a tilt-block basin province (Caminos, 1979a; sequences of continental deposits (fluvial, lacustrine, and
Gordillo and Lencinas, 1979; Lopez et al., 1989). The deltaic). The Cacheuta source rock in the Cuyo basin, for
fluvial, lacustrine, and restricted marine deposits are example, has generated prodigious amounts of oil.
now exposed in the Pampean, Transpampean, and
Petroleum Geology
Famatina ranges.
A magmatic arc with acidic intrusions and extrusions These essentially intracontinental rifts differ from
formed along the western margin of the subcontinent, those of the sub-Andean belt by the large amounts of
what is now the Argentinian–Chilean Frontal Cordillera hydrocarbons they have produced. For example, the
(Groeber et al., 1952; Caminos, 1979b). Cacheuta depocenter of the Cuyo basin complex
The greatest intensity of volcanism occurred during produced over 20 million bbl in 1992. Triassic source
the Late Permian and Early Triassic (Figures 2d, 3, 4). We rocks are widespread: Cuyo-Cacheuta region, Mendoza,
believe that this magmatic arc may explain why some San Juan Precordillera, and intermontane basins such as
70 Urien et al.

Figure 4—Paleogeography of southern South America Figure 5—Paleogeography of southern South America
during the Late Permian–Early Jurassic. Heavy lines are during the Early–Middle Jurassic. Heavy lines are faults.
faults. Legend same as in Figure 3. Legend same as in Figure 3.

Ischigualasto basin. The Cuyo basin (Figure 1) contains rock in the Neuquén basin with its marine faunas and
the most prolific lacustrine bituminous shales of Late turbidite sandstones. This phase of basin subsidence
Triassic age. These are the Cacheuta and Potrerillos come to an end in the Oxfordian with a reorganization of
formations. Estimated ultimate recoverable reserves in stress fields. Old extensional structures were inverted. In
the Cuyo basin are about 1.3 billion bbl of 30˚ API oil and the Neuquén basin, inversion processes formed the
290 bcf of gas. conspicuous east-west trending Huincul arch (Neuquén
dorsal). Figure 6 shows the Late Jurassic–Early Creta-
Early–Middle Jurassic Basins ceous successor basins.
During Middle Jurassic time, a new, predominantly
acidic magmatic event spread lavas and tuffaceous
Basin Development
volcanoclastic materials throughout Patagonia covering
Widespread extension affected the Patagonian crust in large parts of the preceding depocenters. Although the
the Sinemurian, forming a linked system of northwest- sea had withdrawn from much of central Patagonia, the
southeast trending basins (Figure 5). The thickest rift fill Neuquén basin was still inundated (Lesta et al., 1980;
is in the Neuquén basin, which was a marine-influenced Digregorio and Uliana, 1980). A new area of extension
basin throughout much of the Early–Middle Jurassic began in southern Patagonia and the Malvinas plateau
(Digregorio and Uliana, 1980). This tract of extensional (Supko and Perch-Nielsen, 1977; Yrigoyen, 1989a). This
basins extended as far north as the Oriente basin of was the first indication of rifting in the southernmost tip
Ecuador (Groeber et al., 1952; Auboin et al., 1973). of South America and South Africa.
Marine and continental deposits of Early Jurassic age
onlap the basement in these rift basins, from Neuquén Petroleum Geology
basin to the central and southern Patagonia embayments
(Lesta et al., 1980). Important source rock intervals were A noteworthy characteristic of these Early Jurassic
deposited in these basins, such as the Los Molles source extensional basins was their restricted marine circula-
Petroleum Basins of Southern South America: An Overview 71

ARAUCANIAN OROGENY AND


EXTENSION
Late Jurassic –Neocomian Basins
Basin Development
A Late Jurassic (Kimmeridgian) orogenic event
marked the end of the previous pattern of basin
formation and the beginning of a new one. Extensional
faults of the Early Jurassic basins were reactivated in a
reverse sense, resulting in widespread inversion. This
affected basins such as the Neuquén, forming the east-
west Huincul arch, and the Paganzo, where the pull-
apart basins were inverted. This event is known as the
Araucanian orogeny.
After the Araucanian diastrophism, extension reached
a new climax. Subsidence affected the Neuquén basin,
the Central Valley of Chile, western San Jorge basin, and
the Magallanes and Malvinas basins (Charrier and
Vicente, 1972; Stipanicic, 1969a; Yrigoyen, 1989b). For the
first time, Mesozoic extension spread across the older
cratonic interior, affecting basins as diverse as the Orán-
Lomas de Olmedo rift, the Altiplano, and the Paraná
basin. Continental types of deposition dominated.
Along the western margin, a new magmatic arc was
formed as subduction accommodated the Atlantic
opening. The magmatism included andesitic composi-
tions and plutons (Aguirre et al., 1974; González Diaz
and Nullo, 1980; Riccardi and Rolleri, 1980). Basic
volcanism also occurred in the intracratonic basins
(Sprechmann et al., 1981). Some parts of the Cuyo basin
Figure 6—Paleogeography of southern South America and northern and central Patagonia were also affected
during the Late Jurassic–Early Cretaceous. Heavy lines are (Uliana and Biddle, 1988).
faults. Legend same as in Figure 3. Figure 6 summarizes the Late Jurassic–Neocomian
basins along the new Atlantic margin. These basins
involved extensional and strike-slip processes. Sedimen-
tation started with fluvial and lacustrine depositional
tion, which resulted in a diversity of thick deposits of systems and were subsequently subjected to marine
organic-rich marine shales. Organic-rich marls and lime- transgression as extension proceeded (Zambrano, 1980).
stones also occur throughout the stratigraphic column. Examples are the Punta del Este and Salado basins. The
Transgression had flooded several basins, including Colorado basin formed on the grain of the late Paleozoic
Neuquén, San Jorge, the Malvinas plateau, and South Ventana foldbelt (Urien and Zambrano, 1973; Urien et
Malvinas. al., 1981).
In the Neuquén basin, the Pliensbachian–Oxfordian The Valdés-Rawson basin complex is believed to have
Los Molles and Chacay Melehue formations are rich in formed on a volcanic basement, similar to the Colorado
marine black shales and limestones. In some areas, this and San Jorge basins, and was filled with Late Jurassic
organic matter has been matured to a gas-producing and Cretaceous synrift alluvial and lacustrine deposits
stage. These Jurassic source rocks are an important (Zambrano and Urien, 1974; Yrigoyen, 1989a). Farther
reason for the basin’s prospectivity. (Lower Cretaceous south, the San Julián and North Malvinas basins underlie
source rocks are discussed later.) the continental shelf. Because they are part of the process
The NNW-SSE trending rift system of central of Gondwana extension and fragmentation, their sedi-
Patagonia accumulated a thick succession of lacustrine mentary fills can be inferred. We believe that they
bituminous shales in the Oxfordian. These are the contain alluvial sandstones and lacustrine shales similar
Cañadón Asfalto and Aguada Bandera formations. In the to the Rawson and eastern San Jorge basins. East of the
Malvinas plateau and the South Malvinas basin, Patagonian continental margin, fault-bounded half-
Oxfordian marine organic-rich shales have been encoun- grabens are overlain by a passive margin terrace wedge.
tered by JOIDES drilling (Supko and Perch-Nielsen, Along the Andean margin, a new marine transgres-
1977). Recent exploration drilling suggests that these sion was more widespread than those of the previous
potential source rocks may be widespread and extend Jurassic. This trangressive sequence is preserved in the
eastward. Cordillera of Chile and Argentina (Groeber et al., 1952;
72 Urien et al.

Yrigoyen, 1979). These transgressive–regressive


sequences attain their maximum development in the
Neuquén, Magallanes, and Malvinas basins. Further-
more, marine highstand sequences were also widespread
in Neocomian time, such as the deltaic and marine shales
of the Katterfeld Formation in the western San Jorge
basin (Scasso, 1989). The progressive flooding that
started along the Pacific margin in Tithonian time had
reached the eastern edge of the Patagonian platform by
the Barremian (Urien et al., 1981; Zambrano, 1982). This
was probably the route by which the newly evolving
Atlantic rifts were flooded.
Volcanism persisted along the Andean margin
(Aguirre et al., 1974; Muñizaga et al., 1984). Andesitic
lavas and pyroclastics were widespread in western
Patagonia up to Neuquén latitudes (González Diaz and
Nullo, 1980). The distribution of this activity was
probably controlled by old structural fabrics that were
reactivated by extensional and transtensional processes.
Basaltic effusions also occurred in the Pampean ranges
(Gordillo and Lencinas, 1979; Zambrano, 1987; Uliana
and Biddle, 1988). The early Mirano tectonic phase of the
Barremian and Aptian ended this tectonosedimentary
and magmatic cycle (Stipanicic, 1969a).
Petroleum Geology
The Tithonian and Neocomian were characterized by
substantial accumulations of oil-prone source rocks and a
favorable burial history. The greatest volumes of source
rocks occur in southern Argentina. We believe that this
was the principal Mesozoic hydrocarbon-generating
period. The most prolific accumulations occur along the Figure 7—Paleogeography of southern South America
western flank of the subcontinent, from the Malvinas during the Middle Cretaceous. Heavy lines are faults.
plateau toward Mendoza, and probably include the Legend same as in Figure 3.
Central Valley of Chile and the western Altiplano of
Bolivia.
In the Neuquén basin, sea level fluctuations generated excellent oil source properties. The organic material is in
intervals with evaporites, such as the Barda Negra, La the oil window, yielding large amounts of hydrocarbons.
Manga, and Auquilco formations. These were followed This source interval has generated estimated ultimate
by the Tordillo, Vaca Muerta, Quintuco, and Agrio recoverable reserves of 1.4 billion bbl of oil and about 22
formations, which are also attributed to widespread tcf of gas. In the Argentinian part of the Magallanes
transgressive–regressive cycles. The shallow, restricted basin, ultimate recoveries are over 850 million bbl of oil
marine setting is reflected in some of the richest source and 14 tcf of gas (Yrigoyen, 1991a).
rocks of the region. The Vaca Muerta is the principal On the Atlantic margin of southern South America,
source rock in the Neuquén basin (Urien and Zambrano, there are several Late Jurassic–Neocomian rift basins.
1994). The estimated proven and probable reserves in the Lacustrine and shallow marine oil-prone source rocks
Neuquén basin are about 2.3 billion bbl of oil and 18 tcf are anticipated. The primary exploration risk would be
of gas. Oil gravity is 29.5˚–32.2˚ API (Yrigoyen, 1991b). the large quantities of sand shed from the collapsing and
Thick, black, organic shales in the San Jorge basin were eroding Ventana orogeny and potentially poor seals.
deposited in continental environments, including lakes, This tract of basins includes the outer sectors of the
swamps, and estuaries. These are known as the Aguada Salado, Colorado, Rawson, San Julián, and North Mal-
Bandera, Cerro Guadal, and D-129 formations and are the vinas basins, which are still untested.
main rocks of the central and western basins. The
estimated proven and probable hydrocarbon reserves, Middle–Late Cretaceous Basins
which accumulated mainly in lenticular sandstone reser-
voirs of Cretaceous age, are 3.0 billion bbl of 23˚ API oil Basin Setting
and about 4 tcf of gas (Yrigoyen, 1991b).
In the Magallanes basin, widespread sequences of The Aptian brought to a close the Early Cretaceous
continental and marine bituminous shales are docu- history of marine-influenced sedimentation in the
mented. The Cerro Katterfeld, Río Mayer, Springhill, and Neuquén basin. The transition to a new style of subsi-
Palermo Aike formations have black shale intervals with dence (Figure 7) is marked by evaporites such as the
Petroleum Basins of Southern South America: An Overview 73

Huitrin Formation (Groeber et al., 1952; Digregorio and


Uliana, 1980). After this tectonic phase, the Neuquén
basin experienced progressive subsidence dominated by
alluvial sedimentation. Zambrano (1980) speculates that
subsidence may in fact have been of a regional nature,
spanning from the Neuquén to the Colorado basin. This
suggests postrift thermal subsidence. The Cenomanian is
marked by a major unconformity in the Neuquén basin,
reflecting a period of inversion. This event was followed
by red bed accumulation in the Neuquén basin (Urien et
al., 1981).
Alluvial and red bed deposition was ubiquitous,
occurring throughout the Bolsones, Pampas (Laboulaye,
Macachín, and Rosario), San Jorge, Salado, Colorado,
and Rawson basins (Sprechmann et al., 1981; Urien et al.,
1981; Zambrano, 1987). These sedimentary assemblages
comprise reservoirs and seals that were sourced from
earlier lacustrine source rocks.
Marine circulation persisted in the Magallanes and
Malvinas basins with the local development of muddy
calcareous platforms along the rim of the basin (Riccardi
and Rolleri, 1980). The Colorado, Salado, and Rawson
basins coalesced in a broad area of subsidence beneath
the present Atlantic continental slope.
The Andean batholith was probably emplaced in the
Cordillera of southern Patagonia during the Barremian–
Aptian Mirano deformation (Aguirre et al., 1974; Ramos,
1979; Ramos et al., 1982; Muñizaga et al., 1984). This
intrusion was associated with mountain building that
formed a barrier to further marine encroachment from
the Pacific. Magmatic intrusions are also documented in
the coastal Cordillera of Chile. Andesitic compositions Figure 8—Paleogeography of southern South America
and pyroclastics are common in west-central and during the Late Cretaceous–early Tertiary. Heavy lines are
northern Patagonia, as well as in southern Mendoza faults. Legend same as in Figure 3.
(Aguirre et al., 1974; Mingramm et al., 1980)
Although widespread, there was a pattern to this
magmatism. It was generally associated with tectonism
the flexural foreland basin (Zambrano, 1987). This
and transtensional reactivation of old structures. These
shallow foreland basin was covered by a restricted and
intrusive and extrusive igneous rocks appear to be
discontinuous sea from the Pampean region to the
related to strike-slip processes rather than compres-
Eastern Cordillera of Bolivia (Zambrano, 1987). Clastic
sional tectonics. This deformation is expressed in an
depositional systems predominated, including some
angular unconformity in the main Cordillera of
prominent calcareous and evaporitic sequences such as
Mendoza (Polanski, 1964). Marine incursions locally
the Salta and upper Puca groups (Mingramm et al., 1980;
exploited breaks in the tectonic relief of the northern
Cherroni Mendieta, 1977).
Patagonian Cordillera (Auboin et al., 1973; Vicente et
The northern part of Patagonia, including the
al., 1973).
Neuquén basin and the Patagonian platform, were trans-
gressed by a sea that linked the Atlantic with the Pacific,
leaving only the northern Patagonian massif as an
ANDEAN DEFORMATION AND emergent highland (Franchi et al., 1984). Terrigenous
FORELAND BASINS clastics and subordinate carbonates were deposited
(Figure 8).
Intrusive volcanism occurred in the coastal Cordillera
Early Sub-Andean Basins of Chile, in the main and northern Patagonian
cordilleras, and in the sub-Andean belt of central
Basin Setting Patagonia (Groeber et al., 1952; Vergara and Drake, 1979;
The first phase of mountain building, representing the Lesta et al., 1980). There was Late Cretaceous–early
start of Andean deformation, is dated to Late Creta- Tertiary andesitic volcanism in the Andean and sub-
ceous–Paleocene time. This is broadly contemporaneous Andean ranges between 30˚ and 40˚ S lat. Several tectonic
with Laramide deformation of North America. episodes and corresponding magmatic events occurred
Encroachment of these early thrust belt loads initiated in the Andean and sub-Andean ranges. Syntectonic
74 Urien et al.

transgressive-regressive cycles also occurred during the unconformity (Rolleri and Fernandez Garrasino, 1979).
early Maastrichtian, Paleocene, and late Eocene. The Similar depositional settings characterized the middle
earliest Oligocene Incaic movements ended this period Tertiary basins of Patagonia.
(Yrigoyen, 1979). Several compressive phases have been documented.
Three contractional episodes were associated with the
Petroleum Geology Miocene Pehuenche phase alone (Groeber et al., 1952;
Upper Cretaceous–Paleocene limestones and shales of Yrigoyen, 1979). Compression was also associated with
the Salta Group are believed to have generated oils in the tilt block tectonics in the Pampean ranges and with
Orán-Metán basin complex of northwestern Argentina increased magmatic activity. Intrusions in the main
(Figure 1). Caimancito, Martinez del Tineo, and Puesto Cordillera and the Pampean Cordillera have been dated
Guardian are examples of these oil fields. Black shales in to the Oligocene and early Miocene. Other volcanic
the Olmedo Formation and marls and limestones in the centers were the Puna, North Patagonia Cordillera,
Yacoraite Formation are the principal source rocks in the Neuquén basin, and North Patagonian massif.
Orán and Metán basins. The proven and probable
reserves in the Cretaceous Orán basin (= Lomas de
Olmedo basin) are 116 million bbl of 44˚ API oil and 220 Sub-Andean Belt and Foreland Basin
bcf of gas (Yrigoyen, 1991b). The climax of Andean deformation and development
The Ñirihuau basin of west-central Patagonia contains of the modern fold and thrust belt occurred in the middle
Paleocene shallow marine and lagoonal shales, as does Miocene–Pliocene. At this time, the sub-Andean belt or
the neighboring Navidad depocenter of Chile. There are foothills developed by forward propagation. The
organic-rich strata and seeps in the marine Salamanca, Andean orogenesis deepened the foreland basin consid-
Paso del Sapo, and Ñirihuau formations. Similar occur- erably. Old structures beneath the foreland cover were
rences in some Neuquén and Mendoza localities may reactivated, and new structures formed in the foothills.
reflect migration from older units. Tertiary tar deposits This was an important period of hydrocarbon trap
and oil seeps are common, but are generally believed to formation. Some of the best examples are found in the
have been sourced from pre-Tertiary stratigraphy, Chaco basin of Bolivia. Impressive foreland deformation
including San Jorge, Mendoza, northwestern Argentina includes the tilt block province of the Sierras Pampeanas
(Ischigualasto basin), and the Andean margin of the and the Izozog arch of Bolivia. The latter caused uplift
Magallanes basin. Many of these seeps appear to be and erosion of much of the Paleozoic and Mesozoic
related to old strike-slip fault zones. cover. Alluvial fan and fluvial processes dominated the
The Atlantic passive margin and the Malvinas basin tilt block basins (Uliana and Biddle, 1988).
contain Tertiary successions that are more than 3000 m The Pampean and Chaco plains underwent a transgres-
thick. In South Malvinas, flysch and thermal conditions sion that covered the greater part of their surface, followed
are favorable hydrocarbon indicators. The Ciclón x-1 by a gradual regression during the late Miocene–Pliocene.
well, near the Fagnano-Burdwood structural trend, Subsequently, during the late Pliocene–Quaternary, fluvial
shows the presence of hydrocarbons. Reservoirs may be and eolian deposits accumulated.
the risk. Turbiditic and deltaic depositional systems in This middle Miocene phase of Andean orogenesis is
the passive margin terrace wedge are untested. known as the Quechua event. The present topographic
expression of the Andean and sub-Andean ranges, the
Paleogene Andean Basins
Sierras Pampeanas, and the frontal Cordillera were
In the middle Paleogene (late Eocene–Oligocene), established at this time (Mingramm et al., 1980; Uliana
compressive deformation inverted many of the pre- and Biddle, 1988). Uplift of the old volcanic massifs
existing extensional basins and initiated the fold and established new volcanic centers.
thrust belt of the modern Andean ranges. Compressive
deformation also reactivated many preexisting structures
in a strike-slip sense, depending on their relative orienta-
tion. By the middle Miocene, a tilt block province of CONCLUSIONS
rotated blocks and deep basins formed in the
Precordillera and Sierras Pampeanas (Cuerda and The sedimentary basins of the southern part of South
Furque, 1979; Fielding and Jordan, 1988). These processes America preserve a long and complex record of linked
formed the Precordillera as we known it today. basin evolution. Since their inception in the early
Paleozoic by reactivation of inherited basement fabrics,
Middle Tertiary each stage of basin development has reactivated its
Fluvial and eolian landscapes dominated the early predecessor in a new way. These basins are composite in
Neogene Pampean plains. In the Puna and Bolivian structural architecture as well as stratigraphically. They
Altiplano, subsidence was rampant. Terrigenous clastic have a long history of repeated inundation by restricted
sediments, basic lavas, and pyroclastic material were seas that were at times anoxic, progradation of reservoir-
deposited in these basins. The Cuyo foreland basin of rich depositional systems, and recurrent structural
west-central Argentina consists of a westward-thick- episodes. Several basins are prolific oil producers, while
ening wedge of eolian sandstones above an angular others remain largely untested.
Petroleum Basins of Southern South America: An Overview 75

Acknowledgments This contribution is based on the results de Geologia Regional Argentina: Academia Nacional de
of many years of fieldwork and geologic surface and subsurface Ciencias, Cordoba, v. 1, p. 837–869.
observations sponsored by Urien and Associates, whose Cuerda, A. J., and G. Furque, 1979, Precordillera de La Rioja,
financial support is gratefully acknowledged. The authors are San Juan y Mendoza, in J. C. M. Turner, ed., II Simposio de
Geologia Regional Argentina: Academia Nacional de
indebted to A. J. Tankard for his invitation to participate in this
Ciencias, Cordoba, v. 1, p. 455–522.
project and especially for his critical review and correction of De Giusto, J. M., C. A. Dipersia, and E. Pezzi, 1980,
the manuscript. Recognition is likewise due to González Nesocraton del Deseado, in J. C. M. Turner, ed., II
Upton, P. Cazenave, B. Fos, and M. Virasoro for their assis- Simposio de Geologia Regional Argentina: Academia
tance in editing, drafting, and layout and to G. Ortíz for Nacional de Ciencias, Cordoba, p. 1389–1430.
typing the finished text. Digregorio, J. H., and M. A. Uliana, 1980, Cuenca Neuquina,
in II Simposio de Geologia Regional Argentina: Academia
Nacional de Ciencias, Cordoba, v. 2, p. 985–1032.
Eyles, N., 1993, Earth’s glacial record and its tectonic setting:
REFERENCES CITED Earth Science Reviews, v. 35, p. 1–248.
Fernandez-Seveso, F., and A. J. Tankard, 1995, Tectonics and
Aceñolaza, F. G., H. Miller, and A. Tosseli, 1982, Las Rocas stratigraphy of the late Paleozoic Paganzo basin, western
Cristalinas de la Sierra de Ancasti, en el contexto de las Argentina, and its regional implications, in A. J. Tankard,
Sierras Pampeanas Septentrionales: Actas V Congreso R. Suarez, and H. J. Welsink, Petroleum basins of South
Latino Americano de Geología, Buenos Aires, v. 1, America: AAPG Memoir 62, this volume.
p. 333–346. Fielding, E. J., and T. E. Jordan, 1988, Active deformation at
Aguirre, L., R. Charrier, J. Davidson, A. Mpodosis, S. Rivano, the boundary between the Precordillera and Sierras
R. Thiele, E. Tidy, M. Vergara, and J. C. Vicente, 1974, Pampeanas, Argentina, and comparison with ancient
Andean magmatism: its paleogeographical and structural Rocky Mountain deformation: GSA Memoir 171,
setting in the central part of the southern Andes p. 143–163.
(30˚–35˚ S): Pacific Geology, no. 8. Forsythe, R., 1982, The Paleozoic to early Mesozoic evolution
Amos, A. J., 1972, Silurian of Argentina, in W. B. N. Berry and of southern South America: a plate tectonic interpretation:
A. J. Boucot, eds., Correlation of the South American Journal Geologic Society of London, v. 139, p. 671–682.
Silurian rocks: GSA Special Paper 133, p. 5–16. Franchi, N. M., F. E. Nullo, E. G. Sepulveda, and M. A. Uliana,
Andreis, R., A. Inguez, J. Lluch, and S. Rodriguez, 1989, 1984, Las sedimentitas terciarias, in V. A. Ramos ed.,
Cuencas Sedimentarias Argentinas: Universidad Nacional Relatorio Geologia y Recursos Naturales de la Provincia de
de Tucumán, p. 265–298. Rio Negro: IX Congreso Geologico Argentino, Buenos
Auboin, J., A. V. Borrello, G. Cecioni, R. Charrier, P. Chotin, J. Aires, p. 215–266.
Frutos, R. Thiele, and J. C. Vicente, 1973, Esquisse paleo- Franchi, M. R., J. Panza, and L. Barrio, 1989, Depositos
geographique et structurale des Andes Meridionales: Triasicos y Jurasicos de la Patagonia Extraandina: Cuencas
Revue de Geographie Physique et Geologie Dynamique, v. Sedimentarias Argentinas, Serie Correlacion Geologica,
15, p. 11–72. Tucuman, v. 6, p. 347–378.
Baldis, B. A. J., and O. L. Bordonaro, 1984, Cambrico y Furque, G., and A. J. Cuerda, 1979, Precordillera de La Rioja,
Ordovicico de la Sierra Chica de Zonda y Cerro Pedernal, San Juan y Mendoza, in II Simposio de Geologia Regional
Pcia. de San Juan. Genesis del margen continental en la Argentina: Academia Nacional de Ciencias, Cordoba, v. 1,
Precordillera: Actas IX Congreso Geologico Argentino, p. 455–522.
Buenos Aires, v. IV, p. 190–207. González Bonorino, G., and F. González Bonorino, 1991,
Bracaccini, O. I., 1960, Lineamientos principales de la Precordillera de Cuyo y Cordillera Frontal en el Paleozoico
evolución estructural de la Argentina: Petrotecnia, X, no. 6, temprano: terrenos “bajo sospecha” de ser autoctonos:
Buenos Aires, p. 57–59. Revista Geologica de Chile, v. 18, p. 97–107.
Brito Neves, B. B. de, and U. G. Cordani, 1991, Tectonic González Diaz, E., and F. Nullo, 1980, Cordillera Neuquina, in
evolution of South America during the late Proterozoic: II Simposio de Geologia Regional Argentina: Academia
Precambrian Research, v. 53, p. 23–40. Nacional de Ciencias, Cordoba, v. 2, p. 1099–1147.
Caminos, R., 1979a, Sierras Pampeanas Noroccidentales. Salta, Gordillo, C. E., and A. N. Lencinas, 1979, Sierras Pampeanas
Tucuman, Catamarca, La Rioja y San Juan, in J. C. M. de Cordoba y San Luis, in II Simposio de Geologia
Turner, ed., II Simposio de Geologia Regional Argentina: Regional Argentina: Academia Nacional de Ciencias,
Academia Nacional de Ciencias, Cordoba, v. 1, p. 225–291. Cordoba, v. 1, p. 577–650.
Caminos, R., 1979b, Cordillera Frontal, in J. C. M. Turner, ed., Groeber, P., P. Stipanicic, and A. Mingramm, 1952,
II Simposio de Geologia Regional Argentina: Academia Mesozoico: geografia de la Republica Argentina: Buenos
Nacional de Ciencias, Cordoba, v. 1, p. 397–453. Aires, Sociedad Argentina de Estudios Geograficos, 541 p.
Charrier, R., and J. C. Vicente, 1972, Laminary and geosyn- Gust, D. A., K. T. Biddle, D. W. Phelps, and M. A. Uliana,
cline Andes: major orogenic phases and synchronical 1985, Associated Middle to Late Jurassic volcanism and
evolution of the central and Austral sectors of the southern extension in southern South America: Tectonophysics, v.
Andes: Conferencia de la Tierra Solida, Simposio sobre 116, p. 223–253.
investigaciones del manto superior con enfasis en America Harrington, H. J., 1956, Argentina, in W. F. Jenks, ed.,
Latina, Buenos Aires, v. II. p. 451–470. Handbook of South American geology: GSA Memoir 65,
Cherroni Mendieta, C., 1977. El sistema Cretacico de la cuenca p. 129–186.
Cretasica Andina: Revista Tecnica Yacimientos Petroliferos Lesta, P. J., R. Ferello, and G. Chebli, 1980, Chubut
Fiscales Bolivianos, La Paz, v. 6, p. 5–46. Extraandino, in J. C. M. Turner, ed., II Simposio de
Criado Roque, P., and G. Ibañez, 1979, Provincia geologica Geologia Regional Argentina: Academia Nacional de
sanrafaelino-pampeana, in J. C. M. Turner, ed., II Simposio Ciencias, Cordoba, v. 2, p. 1307–1387.
76 Urien et al.

Lister, G. S., M. A. Etheridge, and P. A. Symonds, 1986, Cuencas Sedimentarias Argentinas, Serie Correlacion
Detachment faulting and the evolution of passive margins: Geologica, Tucuman, v. 6, p. 395–418.
Geology, v. 14, p. 246–250 Sprechmann, P., J. Bossi, and J. Da Silva, 1981, Cuencas del
Lopez, O., L. Alvarez, R. Andreis, G. Bossi, I. Espejo, F. Jurasico y Cretacico del Uruguay, in W. Volkheimer and E.
Fernandez, L. Legarreta, D. Kokogian, O. Limarino, and H. A. Musacchio, ed., Cuencas Sedimentarias del Jurasico y
Sesarego, 1989, Cuencas Intermontaneas: Cuencas Sedi- Cretacico de America del Sur: Comite Sudamericano del
mentarias Argentinas, Serie Correlacion Geologica, Jurasico y Cretacico, v. 1, p. 45–126.
Tucuman, v. 6, p. 65–77. Stipanicic, P. N., 1969a, El avance en los conocimientos del
Mingramm, A. R. G., A. Russo, A. Pozzo, and L. Cazau, 1980, Jurasico Argentino a partir del esquema Groeber: Revista
Chubut Extraandino, in J. C. M. Turner, ed., II Simposio de Asociacion Geologica Argentina, Buenos Aires, v. 24,
Geologia Regional Argentina: Academia Nacional de p. 377–388.
Ciencias, Cordoba, v. 1, p. 95–137. Stipanicic, P. N., 1969b, Las sucesiones Triasicas Argentinas,
Munizaga, F., F. Herve, and R. Drake, 1984, Geocronologia K- in A. Amos, ed., Gondwana Stratigraphy: UNESCO,
Ar del extremo septentrional del batolito patagonico en la Buenos Aires, p. 1121–1149.
region de los lagos, Chile: Actas IX Congreso Geologico Stipanicic, P. N., and E. Linares, 1975, Catálogo de edades
Argentino, Buenos Aires, v. 3, p. 133–145. radiométricas determinadas para la República Argentina,
Padula, E. L., and A. R. G. Mingramm, 1969, Subsurface I: años 1960–74, Publ. esp. AGA, serie B, no. 3.
Carboniferous beds of the Chaco-Mesopotamian region, Stipanicic, P. N., and F. Rodrigo, 1969, El diastrofismo Jurasico
Argentina, and their relatives in Uruguay and Brazil: en Argentina y Chile: Actas IV Jornadas Geologicas
Estratigrafia del Gondwana, I Simposio sobre Estratigrafia Argentinas, Mendoza, v. 2, p. 353–368.
y Paleontologia del Gondwana, UNESCO, Paris, Suero, T.,1948, Descubrimiento del Paleozoico superior en la
p. 1025–1040. zona extraandina de Chubut: Boletin Informacion Petroleo,
Pankhurst, R. J., C. W. Rapela, R. Caminos, E. Llambias, and Buenos Aires, v. 28, p. 2–20.
C. Parica, 1992, A revised age for the granites of the central Supko, P. R., and K. Perch-Nielsen, 1977, General synthesis of
Somuncura batholith, North Patagonia massif: Journal of central and South Atlantic drilling results, Leg 39, Deep
South American Earth Sciences, v. 5, p. 321–325. Sea Drilling Project, in P. R. Supko and K. Perch-Nielsen,
Pareja, L., F. Vargas, S. R. Suarez, A. R. Ballacon, C. R. eds., Initial Reports of the Deep Sea Drilling Project, U.S.
Carrasco, and A. C. Villarroel, 1978, Mapa Geologico de Government Printing Office, Washington, D.C., v. 39,
Bolivia: Yacimientos Petroliferos Fiscales Bolivianos, p. 1099–1131.
Servicio Geologico Boliviano. Tankard, A. J., M. P. A. Jackson, K. A. Eriksson, D. K. Habday,
Pezzi, E., and M. E. Mozetic, 1989, Cuencas sedimentarias de D. R. Hunter, and W. E. L. Minter, 1982, Crustal evolution
la region Chacoparanense: Cuencas Sedimentarias of southern Africa: New York, Springer-Verlag, 523 p.
Argentinas, Serie Correlacion Geologica, Tucuman, v. 6., Turner, J. C. M., 1972, Silúrico: Actas de las IV Jornadas
p. 65–77. Geológicas Argentinas, v. III, p. 211–224.
Piatnitzky, A., 1933, Retico y Liasico de los valles del rio Ugarte, F., 1966, La Cuenca compuesta carbonifero-jurasica de
Genoa y Tecka y sedimentos continentales de la sierra de la Patagonia meridional: Anales Universidad Patagonica
San Bernardo: Boletin Informacion Petroleo, Buenos Aires, San Juan Bosco, Ciencia y Geologia, C. Rivadavia, v. 1, p.
v. 10, p. 151–183. 37–68.
Pichowiak, S., A. Bahlung, and C. Breitkreuz, 1987, Paleozoic Uliana, M. A., and K. Biddle, 1987, Permian to Late Cenozoic
volcanic and tectonic evolution in northern Chile: Actas X evolution of northern Patagonia, main tectonic events,
Congreso Geologico Argentino, Tucuman, v. 4, p. 302–304. magmatic activity, and depositional trends, in G. D.
Polanski, J., 1964, Descripcion geologica de la Hoja 25a Volcan McKenzie, ed., Gondwana six: structure, tectonics, and
San Jose: Direccion Nacional de Mineria y Geologia, geophysics: American Geophysical Monograph 40,
Buenos Aires, Boletin 98. p. 271–286.
Prozzi, C. R., and M. Rosso, 1990, Pizarras carbonosas en el Uliana, M.A., and K. Biddle, 1988, Mesozoic–Cenozoic paleo-
basamento de San Luis, Argentina: Actas XI Congreso geographic and geodynamic evolution of southern south
Geologico Argentino, v. I, p. 452–455. America: Revista Brasilera de Ciencias, v. 18, p. 172–190.
Ramos, V. A., 1979, El vulcanismo del Cretacico inferior en la Urien, C. M., and J. J. Zambrano, 1973, The geology of the
Cordillera Patagonica: Actas VII Congreso Geologico basins of the Argentinian continental margin and Malvinas
Argentino, Neuquén, v. 1, p. 423–435. Pau, in A. E. M. Nairn and F. G. Stehli, eds., The Ocean
Ramos, V. A., H. Niemayer, J. Skarmeta, and J. Muñoz, 1982, Basins and Margins, the South Atlantic: New York,
Magmatic evolution of the Austral Patagonian Andes. Plenum Press, v. 1, p. 135–169.
Earth Sciences Reviews, v. 18, p. 311–443. Urien, C. M., J. J. Zambrano, and L. R. Martins, 1981, The
Rapela, C. W., and R. J. Pankhust, 1992, The granites of basins of southeastern South America (southern Brazil,
northern Patagonia and the Gastre fault system in relation Uruguay, and eastern Argentina), including the Malvinas
to the break-up of Gondwana: magmatism and the causes Plateau and southern South Atlantic paleogeographic
of continental break-up: GSA Publication 68, p. 209–220. evolution, in W. Volkheimer and E. A. Musacchio, eds.,
Riccardi, A. C., and E. O. Rolleri, 1980, Cordillera Patagonica Cuencas sedimentarias del Jurasico y Cretacico en America
Austral, in J. C. M. Turner, ed., II Simposio de Geologia del Sur: Comite Sudamericano del Jurasico y Cretacico, v.
Regional Argentina: Academia Nacional de Ciencias, 1, p. 45–126.
Cordoba, v. 2, p. 1173–1306. Urien, C. M., and J. J. Zambrano, 1994, Petroleum Systems in
Rolleri, E. O., and C. Fernandez Garrasino, 1979, Comarca Neuquén Basin, Argentina, in L. B. Magoon and W. G.
septentrional de Mendoza, in J. C. M. Turner, ed., II Dow, eds., The petroleum system—from source to trap:
Simposio de Geologia Regional Argentina: Academia AAPG Memoir 60, p. 513–534.
Nacional de Ciencias, Cordoba, v. 1, p. 771–809. Vergara, M. and R. Drake, 1979, Eventos magmaticos pluton-
Scasso, R., 1989, La cuenca sedimentaria del Jurasico superior icos de los Andes de Chile Central: Actas II Congreso
y Cretacico inferior de la region sudoccidental del Chubut: Geologico Chileno, Santiago, v. 1, p. 19–30.
Petroleum Basins of Southern South America: An Overview 77

Vicente, J. C., R. Charrier, J. Davidson, A. Mpodozis, and S. Authors’ Mailing Addresses


Rivano, 1973, La orogenesis subhercinica: fase mayor de la
evolucion paleogeografica y estructural de los Andes C. M. Urien
Argentino-Chilenos: Actas V Congreso Geologico Buenos Aires Technological Institute
Argentino, Buenos Aires, v. 5, p. 81–98. Paraguay 609, Piso 6˚ “L”
Yrigoyen, M. R., 1979, Cordillera Principal: II Simposio de 1057 Buenos Aires
Geologia Regional Argentina, Academia Nacional de Argentina
Ciencias, Cordoba, v. 1, p. 651–694.
Yrigoyen, M. R., 1989a, Cuencas de Rawson y Peninsula de J. J. Zambrano
Valdés: Cuencas Sedimentarias Argentinas, Serie Correla- Regional Ground Water Institute
cion Geologica, Tucuman, v. 6, p. 467–476. Av. Rioja 489 Norte
Yrigoyen, M. R., 1989b, Cuenca de Malvinas: Cuencas Sedi-
5400 San Juan
mentarias Argentinas, Serie Correlacion Geologica,
Tucuman, v. 6, p. 481–492. Argentina
Yrigoyen, M. R., 1991a, Explanatory notes for the energy
resources map of the Circum-Pacific region, southeast M. R. Yrigoyen
quadrant: USGS Map CP-30, 59 p. National Academy of Sciences
Yrigoyen, M. R., 1991b, Hydrocarbon resources of Argentina: Zavalía 2048, Piso 5˚
Petrotechia, Argentina Petroleum Institute, Buenos Aires, 1428 Buenos Aires
V. XXXIII, p. 38–54. Argentina
Zambrano, J. J., and C. M. Urien, 1974, Pre-Cretaceous basins
in the Argentine continental shelf, in C. A. Burk and C. L.
Drake, eds., The Geology of Continental Margins: New
York, Springer-Verlag, p. 463–470.
Zambrano, J. J., 1980, Comarca de la Cuenca Cretácica del
Colorado en II Simposio de Geología Regional Argentina:
Academia Nacional de Ciencias, Córdoba, v. II,
p. 1033–1070.
Zambrano, J. J., 1982, Posibilidades de Exploracion de hidro-
carburos en la region precordillerana de San Juan: I
Congreso Nacional de Hidrocarburos, Buenos Aires.
Zambrano, J. J., 1987, La estructura superficial y profunda en
el valle de Tulúm, Pcia. de San Juan, Argentina: X
Congreso Geológico Argentino, San Miguel de Tucumán,
v. I, p. 267–270.
Tectonic Subsidence Analysis and Paleozoic
Paleogeography of Gondwana

Kenneth E. Williams
Texaco Inc.
Houston, Texas, U.S.A.

Abstract

C alibrated geohistory analyses of single data points and cross sections are used to construct tectonic
subsidence curves. There are a limited number of distinct curve types, of which seven end-members are
discussed. These curve types and their shapes are genetically related to plate tectonic processes. One regional
backstripped cross section across South America is discussed. The tectonic subsidence curves are used as cali-
bration for a paleogeographic analysis of Gondwana in the Phanerozoic. It is suggested that the southern part
of South America and parts of Antarctica are composed of terranes that were formerly located west of the
present position of the Arequipa massif in Chile and Peru. These displaced terranes formed the western edge
of Gondwana from the Cambrian–Ordovician to the end of the Devonian. During these periods, they were
involved in intracratonic rifting. During the time of the basal Carboniferous unconformity, the terranes were
translated southward to approximately their present positions. Seven reconstructions are presented that are
representative of the major episodes in the development of Gondwana.

Resumen

L os análisis calibrados de geohistoria de los perfiles y puntos de datos simples, se usan para construir de
hundimiento tectónico. Existe una cantidad limitada de tipos distintos de curvas de los cuales comen-
tamos siete miembros de extremo. Estos tipos de curvas y sus formas están genéticamente relacionados a los
procesos de placas tectónicas. Se comenta un perfil regional retrodespejado a través de Sudamérica. Las curvas
de hundimiento tectónico se usan como calibración para un análisis paleogeográfico de Gondwana en el
Fanerozoico. Se sugiere que la parte meridonal de Sudamérica y partes de Antártida están compuestas de
terrenos que estaban situados al oeste de la posición actual del macizo Arequipa en Chile y Perú. Estos terrenos
desplazados formaron el borde occidente de Gondwana del Cambrico-Ordoviciano hasta el final del Devónico.
Durante estos periodos, estuvieron implicados en un agrietamiento intracratónico. Durante el tiempo de la
disconformidad del Carbónico basal, los terrenos se transladaron hacia el sur adoptando aproximadamente sus
posiciones actuales. Se presentan siete reconstrucciones. Los mismos representan los episodios principales en el
desarrollo del Gondwana.

INTRODUCTION deposited during this interval in several basins in South


America. Changing patterns of subduction along the
In this summary of the Phanerozoic history of early Pacific margin are reflected in several regional
Gondwana, special emphasis is placed on the reconstruc- unconformities that bracket the principal geologic
tion of the geologic history and paleogeography of episodes of South America.
southern South America and the surrounding areas. The During the Paleozoic, Gondwana was in a polar
temporal development of the Gondwana margin is also position. A southern ice cap is present on all of the
examined. Paleozoic reconstructions. The maps are shown on a
The Patagonian microplate initially failed to rift from stereographic south polar projection based on the Paleo-
its original position adjacent to the Arequipa massif geographic Atlas Project (Scotese et al., 1979). The recon-
during the Ordovician. It was eventually sheared away structions were tested against a set of tectonic subsidence
in the Early Carboniferous and translated southward and geohistory plots and a set of regional cross sections.
toward its present position. Compression firmly sutured There appear to be a limited number of distinct tectonic
this block onto the continent by the end of Permian time. subsidence curves, each of which is representative of a
A series of seven reconstructions are presented that different tectonic style. These curves were used to
summarize the major episodes in the development of calibrate and test the consistency of the paleogeographic
Gondwana. Source, seal, and reservoir rocks were data during construction of the maps.

Williams, K. E., 1995, Tectonic subsidence analysis and Paleozoic paleogeography of 79


Gondwana, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 79–100.
80 Williams

TECTONIC SUBSIDENCE ANALYSIS


AND DATA POINTS

Tectonic Subsidence Curves


Figure 1 shows a typical tectonic subsidence and
geohistory plot. The data used in the analysis were
derived from the geology of the area, in this example,
from the Serranias Chiquitanas (zona Robore) of eastern
Bolivia (Yacimentos Petroliferos Fiscales Bolivianos,
1972). The location of the dataset is shown as point 6 on
Figure 2. Ages, thicknesses, lithologies, maturity of
organic matter, and estimates of the duration and
amount of erosion were used. In this example, modeling
of the Devonian units assumes that they are within the
oil generative window at present. The data were back-
stripped using the methods described by Sclater and
Christie (1980). Each tectonic subsidence curve was cali-
brated to as much of the available information as
possible. The most common calibrator is vitrinite
reflectance (Cole et al., 1987). This and other indicators of
organic maturity preserve a record of the maximum
temperature to which the rock was subjected and, by
inference, the maximum depth of burial of the sediment
where paleogeothermal gradients can be inferred or
determined (Tissot et al., 1988). Fission track analysis
(Gleadow et al., 1983) was also important since it can
often indicate the time at which the sediments last cooled
below 100˚C. The fully calibrated geohistory analysis
was also compared with adjacent points and distant
points in the same and nearby basins to ensure the
consistency of the interpretation. This was best done on Figure 1—Single data point analysis: (a) tectonic subsi-
reconstructed cross sections based on the backstripped dence data and (b) geohistory analysis. Complete model of
a dataset is based on a calibrated geohistory analysis and
geohistory data (Beaumont et al., 1988). incorporates all available information from nearby datasets
as well as vitrinite reflectance and fission track analyses.
The dataset is from the Sierras Chiquitanas, point 6 on
Tectonic Subsidence Curve Analysis Figure 2. U, unconformity.
Tectonic subsidence curves are sensitive indicators of
the processes that cause basins to subside. They have Voorhoeve and Houseman, 1988). Multiple iterations of
been corrected for sediment loading, paleowater depth, the analysis are generally required in the calibration of
and sediment type and were calibrated on the basis of the curves. Lerche (1990) describes a more mathemati-
geohistory analysis. Comparison of the curves directly cally rigorous technique. These two rift types are the
addresses the questions about what forces were active to result of lithospheric extension. Their analysis may be
cause the basin to subside in its unique way. Figure 3 complex depending on the quality of the data (Karner
shows a classification of tectonic subsidence curves. and Dewey, 1986).
There are seven basic types. Fast flexure is the typical response of a foreland basin
The open rift and closed rift types are the familiar curve to an applied load (Allen et al., 1986; Beaumont et al.,
shapes described by McKenzie (1978). The open rift type 1988). An unconformity at the base of the sediment
characterizes continental borderlands such as the package may mark translation and erosion of the
Atlantic margin of the United States (Watts and Thorne, forebulge (Quinlan and Beaumont, 1984). Examples of
1984) and the west African margin (Brun and Lucazeau, this foreland basin curve type are common worldwide.
1988). The closed rift subsidence curve type is similar, The North Alpine basin (Allen et al., 1986) and the
but is generally intracratonic, such as a failed rift (Burke, Bolivian Chaco basin of Tertiary age are examples.
1977). Examples of this type include the early Paleozoic Distance from the tectonic load is reflected in decreasing
subsidence history of the Anadarko basin of North magnitude of subsidence and the shape of the curve. Far
America (Feinstein, 1981) and the early Paleozoic Chaco from the overthrust load, the curve resembles the slow
basin of Bolivia (Figure 1). A calculated beta (β) flexure or intracratonic curve form. With increased
stretching factor (Watts, 1981) and an extrapolated paleo- distance, the effects of more distant stresses are seen. The
temperature profile facilitate the calibration of the local characteristics of a stress regime become more
geohistory plots (Lucazeau and LeDouaran, 1984; diffuse and less distinctive depending on the distance to
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 81

Figure 2—Index map and reconstruction of Gondwana at Figure 3—Typical tectonic subsidence curve types from
340 Ma. The location of ten data points (stars) and one various tectonic settings. Curve types are indicative of the
cross section (line through point 7) are indicated. tectonic settings of the basins.

the stress source and the lithospheric response to that curve, which is characterized by uplift and erosion. The
externally applied stress. amount of uplift and erosion is determined by the
The slow flexure response has been described as expo- regional geology and by calibration of the geohistory
nentially positive (Thorne and Watts, 1989). It is com- charts (Waples et al., 1992).
monly associated with strike-slip faulting of varying In areas where the magnitude of the subsidence
magnitudes, depending on proximity to the driving curves is relatively small, the effects of parameters other
force. Examples of the lithospheric response to external than the major plate-driving forces can sometimes be
stresses include the Los Angeles basin (Mayer, 1987) and discerned. Lithospheric buoyancy that results from
the Paraná basin. Strike-slip faulting cannot be demon- rheologic differences, as well as plate-bending stresses,
strated in all basins where this curve type is seen. This create compressional or tensional forces that are
curve type indicates only that the stress regime acting on modified locally by the curvature of the earth (Dallmus,
the point was influenced by geographic, time, and 1958). Deviatoric stresses are locally important (Dewey,
strength variations. The patterns of subsidence and uplift 1988).
change depending on how the anisotropic lithosphere A pattern of rapid initial subsidence followed by a
responds to different directions and magnitudes of exter- longer period of exponentially decreasing subsidence
nally sourced stresses. rates, commonly referred to as rift-drift subsidence, is seen
The cratonic response to tectonic subsidence reflects a in open rift basins along continental margins (Watts and
weak stress regime and distant forces. Many interior Steckler, 1981). Closed rift basins form in continental
basins have subsided intermittently in a series of small interiors as a result of deviatoric stresses or by other
pulses of subsidence. This is called the intracratonic curve processes of lithospheric extension. The fast flexure of
type. The shape of each pulse resembles the rift pattern foreland basins changes gradually to slow flexure and
on a small scale. Its duration is generally about 30 m.y., ultimately to a cratonic pattern as the distance from the
with initial rapid subsidence of 1–4 m.y. duration. This is driving force is increased.
characteristic of third-order tectonic cycles (Lambeck et
al., 1987; Cloetingh, 1988). Examples of the intracratonic Data Point Analyses
curve are the Williston basin (Quinlan, 1987) and
Michigan basin (Marshak, 1986) of the central United Several selected data points and one regional cross
States. A second type of cratonic response is a stasis section were chosen to illustrate and constrain the paleo-
82 Williams

Figure 4—Data point analysis for the Bolivian Andes and Figure 5—Data point analysis for the Paraná basin, Brazil.
the Anadarko basin in the United States. (a) Depositional (a) Depositional pattern. See Figure 1 for curve labels. (b)
pattern. See Figure 1 for curve labels. (b) Basement subsi- Basement subsidence rate. Numbers as in Figure 4;
dence rate. Circled numbers in Figures 4–9 refer to the location shown in Figure 2.
ages of the paleogeographic maps: 1, Cambrian; 2,
Silurian; 3, Devonian; 4, Early Carboniferous; 5, Carbonif-
erous; 6, Permian; and 7, Triassic–Jurassic. The upper
dashed line on this plot is the tectonic subsidence curve deposits (Helwig, 1972; Castanos and Rodrigo, 1980).
for the Anadarko deep basin of southern Oklahoma. Permian sedimentation continued at a slower rate and
Location shown in Figure 2. ended with a marine transgression. The Bolivian region
was little affected by the Triassic and Jurassic events of
the Andean belt. There are few sedimentary rocks of this
geographic analysis of Gondwana. The locations of the age in the central part of the basin, although they may
presented data points are shown in Figure 2. The strati- have been stripped by Early Cretaceous erosion, which is
graphic section requires at least age, thickness, and marked by a major unconformity. Subsidence has
lithology information for a minimum analysis. continued episodically to the present. A major increase in
Figure 4 summarizes the dynamics of Phanerozoic the rate of subsidence in the late Tertiary coincided with
basin evolution in Bolivia. Subsidence began in the renewed uplift of the Andes (Lohmann, 1970; Urien et
Cambrian and continued at a rapid rate into the Ordovi- al., 1981; Cobbing, 1985).
cian, when more than 7000 m of clastic sediments were Compared with the Bolivian data, the Paraná basin
deposited. A stretching factor of β = 2.0 was required to (Figure 5) followed a significantly different subsidence
model this rift. The initial mechanical subsidence and and depositional pattern (Northfleet et al., 1969; Fulfaro
subsequent thermal equilibration provided the accom- et al., 1982; Goulart and Jardim, 1982). The early
modation space for the observed sediment fill (McKenzie, Paleozoic subsidence that was so significant in the
1978). The top dashed line in Figure 4 is a comparative Bolivian basin had little or no effect in the Paraná.
tectonic subsidence curve for the Anadarko deep basin of Silurian glaciation left a thin cover of mainly continental
southern Oklahoma (Coffman, 1988). Both the timing rocks. In the Paraná basin, the Devonian was a time of
and magnitude of the two basins are similar. There may marine deposition and source rock accumulation.
in fact have been a tectonic relationship between the two Compared with the vigorous subsidence in the Bolivian
regions based on these similarities and their apparent basin, Paraná subsidence was relatively slight. This is
proximity at that time (Ronov et al., 1984). Subsidence reflected in the Devonian stratigraphic columns: 3200 m
continued at a slower rate in the Silurian due to thermal in Bolivia compared with 1050 m in the Paraná basin.
subsidence. In the Devonian, a new episode of subsi- However, post-Devonian regional erosion was as signifi-
dence and sedimentation affected the basin (Isaacson, cant. This unconformity implies a widespread tectonic
1975), coinciding with regional tectonism and uplift west event (Miliani, 1992). The Precordillera, Chilenia, and
of the Chaco basin. Patagonian terranes are believed to have been translated
A substantial erosional hiatus characterizes the Early southward to their present locations during this time.
Carboniferous. It was during this time that the major During the Carboniferous, the Paraná basin was
terranes west of the basin were translated southward subjected to greater subsidence and was inundated with
along strike-slip faults to their present positions, as postu- great thicknesses of glaciofluvial and glaciomarine rocks
lated in this paper. Carboniferous sedimentation was (Franca and Potter, 1988). Permian subsidence was much
fundamentally different from that of the early Paleozoic greater in the Paraná basin than in the Bolivian Chaco.
(Sempere, 1995). Compared with the underfilled marine Excellent source rocks of the Irati Formation were
basin in the Devonian, the Carboniferous Chaco basin deposited in this interior seaway. Triassic subsidence
was overfilled. The stratigraphy is dominated by glacial was minor and mainly continental (Zalan et al., 1990).
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 83

Figure 6—Data point analysis for the Huallaga Basin, Peru. Figure 7—Data point analysis for the Permian basin, West
(a) Depositional pattern. See Figure 1 for curve labels. (b) Texas, and the eastern Amazon basin, Brazil. (a) Deposi-
Basement subsidence rate. Numbers as in Figure 4; tional pattern. See Figure 1 for curve labels. (b) Basement
location shown in Figure 2. subsidence rate. The upper dashed line is the tectonic
subsidence curve for a data point in the Amazon basin of
Brazil for comparison. Numbers as in Figure 4; location
Figure 6 shows the depositional and subsidence shown in Figure 2.
pattern of the Huallaga basin of Peru (Vargas, 1988),
located on the western flank of the Brazilian shield. The
data point is east of the locus of Cambrian–Ordovician, subjected to increasing subsidence rates. The Triassic was
Devonian, and Early Carboniferous tectonism. In its marked by a significant decrease in the rate of subsi-
early history, subsidence resembled that of the Paraná dence and a change to continental deposition.
basin, with thin lower Paleozoic rocks blanketing a Both of these regions were subjected to similar
basement that apparently had high lithospheric rigidity tectonic processes in the Paleozoic. Both overlie conti-
and low subsidence rates. Slightly higher subsidence nental crust that was attenuated at the end of the
rates occurred in the Permian and Triassic. It was not Precambrian (β = 1.4). Devonian tectonism was not
until the Cretaceous and Tertiary that rapid subsidence significant. In Carboniferous time, North and South
occurred in response to development of the fold and America converged as part of Pangea. This tectonic
thrust mountain belt. episode, and accretion of the Mexican and Central
There are major differences in the appearance of the American microplates west of the two basins, is reflected
curves derived from the Chaco, Paraná, and Huallaga in new and increasing rates of subsidence. Both regions
basins that resulted directly from their relative positions were subjected to similar amounts of stress, although
and the changing tectonic stress fields of South America. neither data point was close to the edge of its respective
Each basin responded differently to the stress fields microplate. The margins of the two microplates
resulting from the same basic tectonic events. Each underwent considerably more tectonism and deforma-
response is due to differences in basement composition tion than at the data points themselves. By Triassic time,
and lithospheric rigidity, distance from the tectonic the stress regimes in the two basins had decreased in
driving force, different stress orientations relative to the magnitude.
crustal fabrics, and different sedimentary sources. The depositional and subsidence data from South
A data point analysis from the Permian basin of West Africa (Figure 8) illustrates development of the Cape and
Texas (Matchus and Jones, 1984) is shown in Figure 7. Karoo basins (Dingle et al., 1983). The subsidence event
This basin had a passive margin setting throughout most that initiated the development of the early Paleozoic
of the Paleozoic. A curve for the Amazon basin of Brazil Cape basin reactivated Eocambrian Pan-African fabrics
(Mosmann et al., 1986) is included for comparison (top (Tankard et al., 1995). The entire lower Paleozoic section
dashed line in Figure 7). The geologic histories of these was deposited along an extensional trailing edge margin.
two seemingly unrelated regions of the world are A major unconformity at the end of the Devonian was
surprisingly similar. Both were facing the Pacific Ocean the result of uplift and erosion of significant amounts of
during most of the Paleozoic and have similar patterns of the pre-Karoo section. Ramos (1986) correlates this latest
subsidence. About 2100 m of sediment accumulated in Devonian diastrophism with accretion of the Chilenia
each basin from Cambrian–Ordovician to Carboniferous terrane in southern South America. This is also the time
time. The curves confirm that each basin was on a when the Malvinas (Falkland) Plateau (believed to have
trailing edge, thermally subsiding margin during this been part of the Patagonian microplate) was accreted to
time. Subsidence was slow and decreased with time. In the edge of the South African Kaapvaal craton and
the middle Carboniferous–Permian, both basins were provided a provenance for Karoo foreland basin sedi-
84 Williams

Figure 8—Data point analysis for southern Africa. (a) Depo- Figure 9—Data point analysis for two locations in Antarc-
sitional pattern. See Figure 1 for curve labels. (b) Basement tica. (a) Depositional pattern. See Figure 1 for curve labels.
subsidence rate. Numbers as in Figure 4; location shown (b) Basement subsidence rate. Solid curve, Pensacola
in Figure 2. Mountains; dashed curve, north Transantarctic Mountains.
Numbers as in Figure 4; location shown in Figure 2.

Figure 10—Index map for Argentina, Paraguay, and Brazil showing the location of the regional cross section and positions
of the principal basins and uplifts During the Paleozoic, the Chaco-Tarija, Chaco-Paraná, and Paraná basin complex was
continuous. They were separated by Mesozoic extension.

mentation (DeWit and Ransome, 1992). Dwyka sedimen- extensional processes that resulted from orogenic
tation suggests a highland source area to the south collapse.
(Visser, 1987). Continued compressional tectonics in the Curves from two data points in Antarctica (St. John,
Late Permian and Triassic resulted in the Cape foldbelt 1986; Bradshaw and Webers, 1988) are illustrated in
and its counterpart in South America, the Ventana fold- Figure 9. The same tectonic episodes affected both but in
belt (Ramos, 1986). Triassic subsidence is attributed to different ways depending on their distance from the
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 85

Figure 11—Regional cross section for Silurian–Carbonif- Figure 12—Regional cross section for Permian–Triassic
erous time slices. Location shown in Figure 10. Stars show time slices. Location shown in Figure 10. Stars, top of oil
the top of the oil generative window. generative window; equal signs, depth to 93˚C.

source of the stress and the stress orientation. Three Sequential Cross Sections
major episodes of subsidence are recognized in these
curves. The first occurs only in the Pensacola Mountains The location of a composite cross section for the
where a Cambrian–Ordovician succession indicates Chaco-Paraná latitudes is shown in Figure 10. This cross
extensional subsidence. In the Devonian, tectonic reacti- section (Figures 11, 12, 13) illustrates the regional nature
vation of the Wilson terrane caused an episode of rapid of the unconformities and the stratigraphic relationships.
subsidence. An unconformity or very slow subsidence in It traverses the Chaco basin of northern Argentina,
the Early Carboniferous characterizes this part of Antarc- crosses the Asuncion arch of Paraguay and the Paraná
tica. It is attributed to accretion of the Patagonian terrane basin of southern Brazil, and continues to the Ponta
and possibly some parts of the Antarctic Peninsula onto Grossa arch near the coast. The western part is in
the main Gondwana landmass. Subduction of oceanic northern Argentina, south of the principal depocenter of
crust below the Antarctic continental crust in the the Bolivian Chaco basin. The section is composed of 14
Permian may be the reason for an increased subsidence data points spread over 1500 km. The data include seven
rate. This event affected the marginal northern wells on the section, three outcrops, and four points
Transantarctic Mountains more than the other area. In derived from regional structural and isopach mapping.
both areas, subsidence essentially ceased in the Triassic Where possible, these data points are calibrated to the
with formation of an unconformity. geochemical data, which indicate former depth of burial
86 Williams

unconformity was significant to the basins on either side


of the arch. There is much more section preserved west of
the Asuncion arch than on the east, partially reflecting
original deposition, tectonism, and erosion. The section
crosses the Las Brenas basin of Pezzi and Mozetic (1989),
which has a thick lower Paleozoic section preserved
(point 6 in Figure 10).
Following the major unconformity, the basins
subsided again beneath a varying thickness of Carbonif-
erous and Permian sediments (Figures 11D, 12A). The
Asuncion arch is the major positive feature on this
profile. This arch and a highland east of the coast were
scoured by Carboniferous glaciers (Visser, 1987; Franca
and Potter, 1988). The Paraná basin developed a distinct
separate identity only in the Carboniferous. The
thickness of the Carboniferous–Permian section that was
deposited and then removed by erosion was not directly
measured. The thicknesses presented here are based on
regional analysis and mapping and are partially
constrained by maturity modeling.
The post-Permian unconformity (Figure 12B) was a
major tectonic event that left the Paraná basin isolated as
a separate depocenter. The Asuncion arch was uplifted
and eroded at this time, much more so than the Ponta
Grossa arch. A significant volume of sediment was also
preserved in the Chaco basin. Overlying the Permian
rocks is the Triassic continental succession (Figure 12C).
This part of the stratigraphy contains widespread
eolianite deposits, indicating an arid paleogeography.
The Serra Geral Formation (Figure 13A) is reported to be
the world’s largest accumulation of basaltic flows on
continental crust. These extrusives cover almost all of the
Paraná basin of Brazil and extend into the Chaco-Paraná
basin in the subsurface. By the end of the Cretaceous
(Figure 13C), erosion had stripped much of this volcanic
layer from the Paraná basin. The Chaco and Paraná
basins subsided in the Cretaceous. Figure 13D shows the
present configuration of the profile.
Figure 13—Regional cross section for Jurassic–Recent
time slices. Location shown in Figure 10. Stars, top of oil
generative window; equal signs, depth to 93˚C; plus signs,
base of oil generative window. PALEOGEOGRAPHIC MAPS
The paleogeographic maps presented here (Figures
and locally may constrain the paleogeothermal gradient 14–21) are based on work by Scotese et al. (1979). South
(Malizia et al., 1993). Four points were available for this polar reconstructions were used, with the south pole at
type of calibration along the cross section. the center. Africa was arbitrarily designated as “up” on
Figure 11A shows the modeled configuration along the Gondwana reconstructions. Microplates that did not
this profile at the end of Silurian time (about 408 Ma). The interact with Gondwana were not analyzed. The maps
Asuncion arch in the middle of the section formed the summarize the paleogeography for seven Phanerozoic
hingeline and eastern margin of the Silurian basin. The time slices.
Chaco-Paraná basin and the Chaco basin were yoked
together in the early Paleozoic before Mesozoic extension Ordovician Paleogeography
separated them (Starck, 1995). The early Paleozoic pattern
of deposition where the section thickens dramatically to Figure 14 shows a Gondwana reconstruction for the
the west across the Asuncion arch in the direction of the Ordovician, approximately 475 Ma. Gondwana was the
Bolivian basin continued into the Devonian (Figure 11B). southern continent, with the North American and
This established the basin architectural pattern by the end European landmasses more equatorial. Lash (1987) has
of the first major depositional sequence. Figure 11C described the collision between North America and
(about 340 Ma) shows the structural and stratigraphic Europe. A volcanic arc appears to have been present off
relationships at the pre-Carboniferous unconformity. This the northern margin of Gondwana, the remnants of
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 87

Figure 14—Paleogeographic map for the Ordovician, 475 Ma. Legend is for Figures 14–21. Data for the presence of glaciation
at each time interval came from various sources, principally Caputo and Crowell (1985), Chatterjee and Hotton (1986), St.
John (1986), and Van Houten and Hargraves (1987).
88 Williams

which are now located in the Andes of Colombia (Burgl, limestones and Silurian diamictites) indicate the possi-
1967; Barrett, 1988). A large ice cap is believed to have bility of lateral displacement (Martinez, 1980). This lateral
been located in central Africa based on numerous glacial faulting episode is postulated to have occurred in the
indicators shown on the paleogeographic reconstruction. early Carboniferous.
An oceanward-dipping subduction zone was present Volcanics in the Sierras Pampeanas terrane of Argen-
along the Antarctic and Australian margin (Veevers, tina indicate that extensional activity persisted into the
1980; Kleinschmidt and Tessensohn, 1987). Other authors Silurian (Ramos et al., 1986). Subsidence continued in the
infer the presence of a landward-dipping subduction Bolivian Chaco basin as shown by the tectonic subsi-
zone at this time (St. John, 1986). The Wilson terrane was dence curves. However, the rate was much slower than
accreted onto Antarctica along the Ocloyic-Ross- in the Cambrian–Ordovician, indicating that extension
Delamerian-Adelaide orogenic belt (Figure 14) (Coira et was superseded by a cooling phase. Sand-prone facies
al., 1982; Stump, 1987). are more prevalent on the southwestern margin of the
A large landmass is believed to have existed west of basin than along its northeastern flank (Cobbing, 1985).
present-day Peru and Bolivia in the Ordovician and is This was a continuation of the pattern established in the
shown in Figure 14 as the Patagonian block. It has been Cambrian–Ordovician.
variously described in the literature as the Arequipa
massif (Cobbing, 1985; Isaacson, 1975), Belen and Mejil- Devonian Paleogeography
lones (Ramos, 1986), southeast Pacific paleocontinent
(Dalmayrac, 1988), proto-Puna (Herve et al., 1987), and Devonian marine shales (380 Ma) are the best known
Pampean massif (Helwig, 1972). At that time, there was source rocks in the Bolivian and northern Argentinean
no continental landmass forming the southern tip of producing basins (Illich et al., 1981), and sandstones of
South America (Ramos, 1989; Visser, 1993). Ordovician that age are significant reservoirs. Consequently, the
extension affected the southwestern margin of Devonian has significant commercial interest. Shales of
Gondwana, forming a tract of intracratonic rift basins in Devonian age also form source rocks for hydrocarbons in
Bolivia and northern Argentina (Cobbing, 1985; Dal- the Permian and Anadarko basins of North America
mayrac, 1988; DeWit and Ransome, 1992). Over 7 km of (Schmoker, 1986) and in the Amazon basin of Brazil
sediments accumulated in this rapidly subsiding rift (Mosmann et al., 1986).
system. Rifting was extensive, but not enough to create The south pole was located on the margin of
oceanic crust. Counterclockwise rotation of a portion of Gondwana in the Devonian (Figure 16) (Scotese et al.,
the remaining remnant of the Arequipa massif is 1979). The climate was warmer and the ice caps slightly
recorded as having occurred in latest Cambrian–Early smaller than in the Silurian. Glacial deposits have been
Ordovician time (Forsythe et al., 1993). A small volume reported from the Amazon basin (Mosmann et al., 1986;
of Ordovician volcanics is present in the Bolivian Andes Van Houten and Hargraves, 1987; Veevers and Powell,
(Padula, 1972). The landward margin of the rift system 1987) at a latitude close to 30˚S on the reconstructions of
was probably marked by the Asuncion arch. The Scotese et al. (1979). Glacial diamictites occur in the
regional cross section (Figures 11–13) suggests a Upper Devonian of the Solimoes (western Amazon) and
threefold increase in the thickness of the sedimentary in the Amazon and Parnaiba basins of Brazil. This was
section west of the Asuncion arch in Paraguay. most likely in a temperate climatic zone. The occurrences
This inferred highland west of the Bolivian rift system are in an east-west trending tract of basins and are attrib-
was a source area for the voluminous clastic rocks that uted to Alpine glaciation along their southern margin.
filled the basin (Cobbing, 1985). The Brazilian shield was The Brazilian shield is also believed to have supported
probably not elevated enough to form a substantial glaciers.
source area (Figures 5, 6). To the south, the Pampeanas Colombia was characterized by marine shale deposi-
terrane was the site of major volcanic activity attributed tion in the Devonian and an Appalachian fauna similar
to an accreting island arc (Acenolaza, 1976; Coira et al., to counterparts in North America (Copper, 1977; Barrett,
1982; Ramos et al., 1986; Herve et al., 1987; Wilner et al., 1988; Van Der Voo, 1988). The Gondwana Malvinokaffric
1987) faunal province and the Appalachian faunal province are
believed to have been separated by the same topographic
Middle Silurian Glaciation high that separated the Bolivian and Amazon basins. The
Devonian of South Africa was dominated by deltaic
The Middle Silurian (Figure 15) was a time of wide- deposition (Theron and Loock, 1988). Marginal marine
spread glaciation about 425 Ma (Grahn and Caputo, environments persisted through most of the Devonian.
1992). The distribution of these glacial sediments Devonian rocks deposited in the Chaco and Chaco-
suggests that the south pole was centered in the interior Paraná basins were derived from the Brazilian shield as
of Africa near the center of the landmass (Caputo and well as a western source terrane. Sandstone facies are
Crowell, 1985; Van Houten and Hargraves, 1987; Cocks more abundant along the basin margins (Cobbing, 1985).
and Fortney, 1988). Minor accretion occurred along the The Chaco-Paraná basin of Paraguay contains a largely
Antarctic and Australian margin of Gondwana. This argillaceous fill. In the Chaco basin of Bolivia and north-
tectonic event has been referred to as the Borchgrevink– western Argentina, the western orogen was the primary
Tasman orogen (St. John, 1986). Some Silurian outcrops source of sediment (Mingramm et al., 1972; Isaacson,
in the Andes (e.g., the association of Silurian warm water 1975; Sempere, 1995). In fact, the western source was
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 89

Figure 15—Paleogeographic map for the Middle Silurian, 425 Ma. Legend same as in Figure 14.

considerably more substantial than the present-day south of the Patagonian terrane, is believed to have
Andean belt (Isaacson, 1975). This voluminous sediment moved to its present position along a strike-slip system.
supply was accommodated by a greater rate of subsi- Intrusive and extrusive volcanics are present along the
dence in the Bolivian Chaco basin, as shown by the Sierras Pampeanas and along the western margin of the
tectonic subsidence curves (Figure 4). The present author subsiding Chaco basin (Ramos et al., 1986; Wilner et al.,
believes that this enormous volume of sediments is 1987). Late Devonian uplift also affected the Amazon
explained by a growing orogen augmented by the Pata- basin region. This interpretation implies that there was
gonian microplate which was then located west of no continental crust forming the southern tip of South
northern Chile and Peru (Figure 16). Bird and Molenaar America. Devonian compressional tectonics in northern
(1992) have resorted to a similar interpretation to explain Argentina are referred to as the Chanic orogeny (Coira et
a northerly source of sediments for the Paleozoic North al., 1982; Cobbing, 1985).
Slope of Alaska. An isolated remnant of the Devonian warm water
Figure 16 shows the Patagonian terrane west of the Appalachian fauna is present west of the Arequipa
Peru–Chile coast and west of the Arequipa massif. The massif in southern Peru (Cobbing, 1985). This assem-
Chilenia terrane of Ramos et al. (1986), located just to the blage best matches analogous outcrops in Ecuador. This
90 Williams

Figure 16—Paleogeographic map for the Devonian, 380 Ma. Legend same as in Figure 14.

outcrop appears to be located on one of the many Interaction between the major plates of Pangea
displaced terranes that underwent southward transla- created the Appalachian orogeny of eastern North
tion. America (Hatcher, 1988). Thrust faulting occurred as far
south as the Ouachita Mountains of the Arkoma basin
Early Carboniferous Paleogeography where deformation was relayed to oceanic crust between
North America and Gondwana (Thomas and Schenk,
At the end of the Devonian about 340 Ma, a major 1988). Stresses relayed from the linked oceanic crust also
change in the paleogeography of Gondwana occurred affected the landmass as far away as the Amazon basin
(Figure 17). It is characterized by a widespread unconfor- region (Figure 7). It is likely that the oceanic crust of the
mity. Evidence is abundant that several intracratonic Pacific region changed from passive cooling to active
highland regions developed in South America at this subduction at this time. All tectonic subsidence curves in
time. Translation of the various terranes to their present this paper (Figures 1, 4–9) show a simultaneous uncon-
positions in southern South America and Antarctica also formity. Interactions between the major plates of the
probably occurred at this time. oceanic Pacific region and those of continental
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 91

Figure 17—Paleogeographic map for the base Carboniferous unconformity, 340 Ma. Legend same as in Figure 14.

Gondwana were the driving force for this tectonic episode. Precordillera terrane moved at an early stage because it
Southward translation of the microplates from their was outboard of the large western terrane and accumu-
initial positions west of the Arequipa massif to their lated carbonates on a trailing margin in the Ordovician.
respective locations in Patagonia and Antarctica Extensive deformation affected its western margin after
probably occurred during the Late Devonian and Early emplacement. The Patagonian, Chilenian, and other
Carboniferous. The precise timing is obscured by associated terranes probably moved later. The Paganzo
younger tectonism, deformation, and erosion. However, basin is located on the northern boundary of the
this history is recorded in a major unconformity that Precordilleran terrane. It subsided rapidly in the
spans much of southwestern Gondwana. This is the Namurian (Archangelsky et al., 1987) immediately after
Eohercynian unconformity of some authors (Lopez- accretion of the displaced terranes and may reflect a
Gamundi and Rossello, 1993). period of local extension. Volcanics and inferred
Figure 18 shows an attempt to reconstruct this terrane magmatic arc rocks possibly record accretion of the
translational history, although the order in which these Patagonian–Malvinas terrane complex 340 Ma (Ramos,
microplates moved is speculative. It is probable that the 1986). In this context, the train of volcanics marks the
92 Williams

Pangea was fully assembled at this time (Figure 19).


The Amazon basin subsided rapidly as a result of
convergence of the two supercontinents as well as subsi-
dence along the western margin of the continent
(Mosmann et al., 1986). Low hills fringed the southern
margin of the basin; these were erosional remnants of the
earlier mountain belt that had been the source of glacial
deposits in the basin (Figure 17). There was also signifi-
cant tectonism along the Antarctic–Australian margin as
a result of landward-dipping subduction, reflected in
volcanics and mountain building (St. John, 1986; Veevers
and Powell, 1987; Audley-Charles, 1988).
There is significant commercial oil and gas production
from Carboniferous glaciofluvial rocks in Oman (Levell
et al., 1988) and from the Chaco and Tarija basins of
Bolivia and northern Argentina. Both regions were
located at about 30˚ S lat during the Carboniferous, at the
limit of glacial influence. Based on experience in Bolivia, I
suggest that in both regions the deposits were sourced by
mountain glaciation. The paleogeography of the western
Figure 18—Inferred reorganization of terranes. PB,
Paganzo basin. 1, Precordillera; 2, Patagonia; 3, Chillenia;
Chaco basin consisted of wide glaciofluvial valleys with
4, Parts of Antarctica and other scattered terranes. northward-flowing streams (Helwig, 1972; Tankard et
al., 1995). The depositional landscape consisted of valleys
up to 500 m deep that were filled with stacked glacioflu-
suture zone. The terrane boundary of the northern edge vial complexes, lacustrine shales that were locally
of Patagonia is south of the Sierra de la Ventana and is varved, and glacial diamictites with striated clasts. It was
obscured by younger sedimentary deposits. The absence an extremely sand-prone depositional setting but which
of contemporaneous compressional deformation became more argillaceous to the north (Castanos and
suggests that the left-lateral movement had a transten- Rodrigo, 1980). Glaciers are believed to have existed in
sional sense of behavior. the fold and thrust belt that formed the western flank of
Several highland or mountainous areas owe their the Chaco basin and on the elevated cratonic highlands
uplift to this tectonic episode. Examples include the to the east.
Asuncion arch and the region of the truncated Arequipa There were several marine incursions deep into the
massif (Messner and Wooldridge, 1964; Franca and interior of South American Gondwana (Northfleet et al.,
Potter, 1988). These highlands formed local glacial 1969; Frakes and Crowell, 1972; Soares et al., 1977; Franca
centers at various times during the Carboniferous. and Potter, 1988). The incursions into the Paraná basin of
This interpretation of the Patagonian block is based on southern Brazil are well documented. These seas
Martinez (1980), but it differs from his in that translation probably flooded from the south (Figure 19).
is believed to have occurred over a short period of time Glacial marine and glaciofluvial depositional systems
in the Late Devonian and Early Carboniferous during the influenced the Paraná basin of southern Brazil and
Chanic tectonic episode. The present interpretation northern Argentina (Mingramm et al., 1972; Soares et al.,
involves several small blocks. Besides the left-lateral 1977). In South Africa, the same evidence of massive
scheme emphasized here, there were apparently also glaciation is preserved (Visser, 1987). According to
local right-lateral movements (Acenolaza and Toselli, Dingle et al. (1983), the Cape foldbelt in South Africa and
1976; Ramos, 1989). These opposing senses of displace- its counterpart, the Ventana foldbelt in Argentina, both
ment are not mutually exclusive and are common in began deformation at this time by inversion of older
wrench fault terranes, such as the Garlock fault in the extensional depocenters. This deformation is attributed
Death Valley area of California (King, 1977). to readjustment of the accreted microplates of the Pata-
gonian block.
Carboniferous Paleogeography
The Carboniferous reconstruction (Figure 19) shows Permian Paleogeography
the paleogeography at about 290 Ma for one of the many
glacial maxima that affected Gondwana. That waxing Pangea was influenced by a wide variety of stresses in
and waning of ice sheets occurred on numerous occasions the Permian at about 250 Ma as a precursor to the frag-
is shown by the large number of cyclothems in Kansas mentation that would occur in the Mesozoic (Veevers et
and Oklahoma that have been attributed to glacial al., in press). The south pole was located off Antarctica
eustatic sea level oscillations and the effects of tectonism (Scotese et al., 1979; Caputo and Crowell, 1985) (Figure
(Lopez-Gamundi, 1989; Klein and Kupperman, 1992). 20), which resulted in a warmer climate than in Carbonif-
Gondwana was subjected to continental ice sheets erous–Early Permian time. Glaciation was more
(Veevers and Powell, 1987; Visser, 1987; Eyles et al., 1995). restricted (Veevers and Powell, 1987).
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 93

Figure 19—Paleogeographic map for the Carboniferous, 290 Ma. Legend same as in Figure 14.

A major orogenic episode affected the southwestern Permian glaciation along the margin of Antarctica and
margin of Gondwana in the middle Permian–Triassic Australia deposited sediments in a back-arc basin setting
interval (Lawver and Scotese, 1987). A broad tract of fold that was apparently related to landward-dipping
mountains was created in Australia, Antarctica, southern subduction (Elliot, 1975). The island arc system died out
Africa, and South America by selective inversion of older during the Permian. By the end of the Permian, there
extensional basins. It is believed that there was a left- was an active subduction zone along the northern
lateral component of shearing in South America that margin of Australia (Veevers, 1980; Audley-Charles,
reflects adjustments of the Patagonian terrane (Figure 20). 1988). The Permian was also a time of shrinking glaciers
This strike-slip behavior may also explain the juxtaposi- and climate amelioration.
tion of reefal carbonates and cold water faunas (Helwig, In India, a radiating suite of rift basins was invaded by
1972; Martinez, 1980). Sedimentary basins were yoked a marine transgression (Chatterjee and Hotton, 1986).
together along this tectonic belt (Veevers, 1980). Extrusive This extensional system and the interior rift of Australia’s
and intrusive volcanics were relatively common along the Cooper basin (Wopfner, 1981) reflect the tectonic linkage
continental margin. across Gondwana or the farfield effects of intraplate
94 Williams

Figure 20—Paleogeographic map for the Permian, 250 Ma. Legend same as in Figure 14.

stresses (Cloetingh, 1988). Permian tectonism and volcanism were widespread


Parts of the Bolivian Chaco basin subsided during the along the length of southern South America (Martinez,
Permian, while other parts appear to have been eroded. 1980; Coira et al., 1982; Ramos et al., 1986). A wide band
The shallow Copacabana seaway collected terrigenous of volcaniclastics, basic flows, intrusives, and shearing
clastics and carbonate sediments as well as volcanic occurs north of the Patagonian terrane in southern
detritus (Cobbing, 1985). The sequence thickens toward Argentina (Miller, 1981). The progressive accretion of
the northwest and Peru (Castanos and Rodrigo, 1980). terranes is recorded in the radiometric ages that get
Overall, the subsidence of the Chaco basin was minimal. younger in that direction (Forsythe, 1982; Herve et al.,
In contrast, the Paraná basin subsided continuously 1987).
during the Permian. Several internal arches and local The most promising hydrocarbon source rock in the
depocenters reacted differently to prevailing stress fields. Paraná basin is the Irati Formation (Padula, 1969). It has
In particular, the Asuncion, Ponta Grossa, and Rio Grande been sufficiently deeply buried to be mature over a large
arches were periodically uplifted so that sediments part of the basin (Cerqueira and Neto, 1986). In South
onlapped and blanketed them (Gama et al., 1982). Africa, the equivalent of the Irati formation is the
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 95

Figure 21—Paleogeographic map for the Triassic, 200 Ma. Legend same as in Figure 14.

Whitehill formation of the Karoo basin (Anderson and End of Triassic Paleogeography
McLachlan, 1979). The glacial deposits of the Paraná
Itarare Formation are equivalent to the Dwyka Group of Continental conditions prevailed throughout
South Africa. Gondwana during the Triassic, leaving a widespread
The Cape foldbelt was initiated during the Permian. record of eolianites and redbeds (Figure 21). A large
Cratonward, the flexural Karoo basin subsided in unconformity at about 200 Ma marks the base of the
response to an applied load. Dingle et al. (1983) have Mesozoic over most regions (Padula, 1972; Milani, 1992).
suggested that the Malvinas plateau south of the Karoo Extension affected Patagonia and Africa. Volcanic
basin also formed a highland. Continued compression activity occurred locally. The landmass was farther from
and inversion of the Cape foldbelt is attributed to adjust- the south pole, which explains the warmer climates
ments of the Patagonian terrane resulting from (Scotese et al., 1979).
continued oceanic–continental plate interactions Mountain building by intracratonic inversion
96 Williams

continued into the Triassic. In southern Africa and volcanism, and emplacement of batholiths imply
southern South America, inversion was a selective subduction along the margin of the continent. Organic-
process and affected basins that were already offset rich shales formed in the intracratonic basins (e.g., Irati
across accommodation zones. The result was an irregular shales of Brazil and Whitehill shales of South Africa).
orogen with marked offsets (Tankard et al., 1995). Later Nonmarine conditions prevailed over much of the
Triassic extension by orogenic collapse affected both continent during the Triassic.
sides of the Atlantic. Basins such as the Cuyo in
Patagonia and the Moltero in South Africa were land-
locked and accumulated organic-rich lacustrine source
rocks and braided fluvial deposits shed from the orogen Acknowledgments I thank Glenn Ware and Don Irwin for
(Uliana and Biddle, 1987; Tankard et al., 1995). In Antarc- their support and Oscar Lopez-Gamundi for discussions on
tica, several terranes display strike-slip and transpres- some of the concepts pursued in this paper. Jim Kotchavar
sional movements related to Triassic subduction (Elliot, drafted the paleogeographic maps. Margy Walsh is thanked for
1975; Wilson et al., 1989). her library searches. I also wish to thank Texaco for support
and permission to publish this paper.

SUMMARY
REFERENCES CITED
The southern part of South America and the margin of
Antarctica were apparently built by a mélange of exotic Acenolaza, F. G., 1976, The Ordovician System in Argentina
terranes and microplates that were originally located and Bolivia, in M. G. Bassett, ed., The Ordovician System:
west of the present-day Arequipa massif in Chile. These Proceedings of a Paleontological Association Symposium:
terranes formed the western edge of Gondwana until the Birmingham, September 1974, Cardiff, University of Wales
end of the Devonian. In the early Paleozoic, the region Press, p. 479–487.
Acenolaza, F. G., and A. J. Toselli, 1976, Consideraciones
was affected by intracratonic rifting. Separation and estratigráficas y tectónicas sobre el Paleozoico inferior del
southward translation of the terranes is marked by a noroeste Argentino: Congresso Latinoamericano
conspicuous Late Devonian–Early Carboniferous uncon- Geologico Memoir II, p. 755–763.
formity. The younger Paleozoic paleogeographies are Allen, P. A., P. Homewood, and G. D. Williams, 1986,
very different as a result of the movement of large land- Foreland basins: an introduction, in P. A. Allen and P.
masses along the continental margin. The resulting Homewood, eds., Foreland basins: Special Publication of
tectonism and plate readjustments culminated in the the International Association of Sedimentologists, no. 8,
Permian–Triassic Ventana and Cape foldbelts. The later p. 3–12.
Mesozoic was marked by rifting and the breakup of Anderson, A. M., and I. R. McLachlan, 1979, The Oil-shale
Gondwana. potential of the Early Permian White Band Formation in
Southern Africa: Geological Society of South Africa Special
Ordovician rifting in the Bolivian Chaco basin caused Publication 6, p. 83–89.
the accumulation of more than 7000 m of sediments. A Audley-Charles, M. G., 1988, Evolution of the southern
large percentage of these deposits had its source from a margin of Tethys (north Australian region) from Early
large landmass located west of the present edge of South Permian to Late Cretaceous, in M. G. Audley-Charles and
America. I believe that this landmass consisted of a A. Hallam, eds., Gondwana and Tethys: Geological Society
mosaic of terranes, including those that would eventu- of London, Special Publication 37, p. 79–100.
ally build Patagonia and the Antarctic Peninsula. In Archangelsky, S., A. J. Amos, R. R. Andreis, C. L. Azcuy, C. R.
Silurian time, the Chaco basin continued to subside and Gonzalez, O. Lopez-Gamundi, and N. Sabattini, 1987, El
accumulate sediment from the Patagonian microplate. Systema Carbonifero en la Republica Argentina: 255 p.
Glaciation occurred over large parts of the continent. In Barrett, S. F., 1988, The Devonian System in Colombia, in N. J.
McMillan, A. F. Embry, and D. J. Glass, eds., Devonian of
the Devonian, higher rates of subsidence in the Chaco the World: Canadian Society of Petroleum Geologists
basin are attributed to subduction of oceanic crust to the Memoir 14, v. 1, p. 705–716.
west of the Patagonian microplate. The Devonian Beaumont, C., G. M. Quinlan, and J. Hamilton, 1988, Orogeny
sequence contains prominent source and reservoir rocks and stratigraphy: numerical models of the Paleozoic in the
in the Chaco and Amazon basins. The Late Devonian– eastern interior of North America: Tectonics, v. 7,
Early Carboniferous unconformity represents the time of p. 389–416.
lateral southward translation of the Patagonian, Bird, K. J., and C. M. Molenaar, 1992, North Slope foreland
Precordillera, and other terranes to approximately their basin, Alaska, in R. W. Macqueen and D. A. Leckie, eds.,
present positions. Highlands were created in the interior Foreland basins and fold belts: AAPG Memoir 55,
of Gondwana as a result of intraplate stresses. p. 363–394.
Bradshaw, M. A., and G. F. Webers, 1988, The Devonian rocks
The Carboniferous was characterized by overfilled of Antarctica, in N. J. McMillan, A. F. Embry, and D. J.
sand-prone basins and widespread glaciation (Figure 19). Glass, eds., Devonian of the World: Canadian Society of
Waxing and waning of ice sheets deposited repeated Petroleum Geologists Memoir 14, v. 1, p. 783–795.
cyclothems. On the fringes of the ice sheet, mountain Brun, M. V. L., and F. Lucazeau, 1988, Subsidence, extension
glaciers were an important source of sediments in Oman and thermal history of the west African margin in Senegal:
and Bolivia. In Permian time, mountain building, Earth and Planetary Science Letters, v. 90, p. 204–220.
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 97

Burgl, H., 1967, The orogenesis in the Andean system of Eyles, N., G. Gonzales-Bonorino, A. B. Franca, C. H. Eyles, O.
Colombia: Tectonophysics, v. 4, p. 429–443. Lopez–Paulsen, 1995, Hydrocarbon–bearing late Paleozoic
Burke, K., 1977, Aulacogens and continental breakup: Annual glaciated basins of southern and central South America, in
Reviews of Earth and Planetary Science, v. 5, p. 371–396. A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum
Caputo, M. V., and J. C. Crowell, 1985, Migration of glacial basins of South America: AAPG Memoir 62, this volume.
centers across Gondwana during Paleozoic Era: GSA Feinstein, S., 1981, Subsidence and thermal history of the
Bulletin, v. 96, p. 1020–1036. southern Oklahoma aulacogen: implications for
Castanos, A., and L. A. Rodrigo, 1980, Paleozoico superior de petroleum exploration: AAPG Bulletin, v. 65,
Bolivia: Anais da Academia Brasileira de Ciencas, Revista p. 2521–2533.
Brasileira de Biologia, v. 52, p. 851–864. Forsythe, R., 1982, The late Paleozoic to early Mesozoic
Cerqueira, J. R., and E. V. Neto, 1986, Papel das intrusoes de evolution of southern South America: a plate tectonic
diabásio no processo de geração de hidrocarbonetos na interpretation: Geological Society of London, v. 139,
bacia do Paraná: Third Congresso Brasileiro de Petroleo, p. 671–682.
Rio de Janeiro, p. 1–15. Forsythe, R. D., J. Davidson, C. Mpodozis, and C. Jesinkey,
Chatterjee, S., and N. Hotton III, 1986, The paleoposition of 1993, Lower Paleozoic relative motion of the Arequipa
India: Journal of Southeast Asian Earth Sciences, v. 1, no. block and Gondwana; paleomagnetic evidence from
3, p. 145–189. Sierra de Almeida of northern Chile: Tectonics, v. 12,
Cloetingh, S., 1988, Intraplate stress: a new element in basin p. 219–236.
analysis, in K. L. Kleinspehn and C. Paola, eds., New Franca, A. B., and P. E. Potter, 1988, Estratigrafia, ambiente
perspectives in basin analysis: New York, Springer- deposicional e análise de reservatório do Grupo Itararé
Verlag, p. 205–230. (Permocarbonífero), bacia do Paraná, parte 1: Boletin de
Cobbing, J. E., 1985, The central Andes: Peru and Bolivia, in geociencias da Petrobras, Rio de Janeiro, v. 2, p. 147–191.
A. E. M. Nairn, F. G. Stehli, and S. Uyeda, eds., The ocean Frakes, L. A., and J. C. Crowell, 1972, Late Paleozoic glacial
basins and margins, Vol. 7A, The Pacific Ocean: New geography between the Paraná basin and the Andean
York, Plenum Press, p. 219–264. geosyncline: Anais da Academia Brasileira Ciencias, v. 44,
Cocks, L. R. M., and R. A. Fortney, 1988, Lower Paleozoic p. 139–145.
facies and faunas around Gondwana, in M. G. Audley- Fulfaro, V. J., A. R. Saad, M. V. dos Santos, and R. B. Vianna,
Charles and A. Hallam, eds., Gondwana and Tethys: 1982, Compartimentação e evolução tectonica da bacia do
Geological Society of London Special Publication 37, Paraná, in R. Yoshida and E. Gamma, Jr., eds., Geologia da
p. 183–200. bacia do Paraná, Reavaliação da potencialidade e prospec-
Coffman, W. J., 1988, SW-NE diagramatic cross-section across tividade em hidrocarbonetos: Pauli Petro Consorcio, Sao
Anadarko basin showing structural configuration and Paulo, p. 75–115.
stratigraphy, in R. Bailey, Jr., and N. J. Hyne, eds., Gama, E., Jr., A. N. Bandeira, Jr., and A. B. Franca, 1982,
Petroleum geology of the mid-continent: Tulsa Geological Distribuição especial e temporal das unidades litoestrati-
Society Special Publication 3, p. 143. gráficas Paleozoícas na parte central da bacia do Paraná,
Coira, B., J. Davidson, C. Mpodizis, and V. Ramos, 1982, in R. Yoshida and E. Gamma, Jr., eds., Geologia da bacia
Tectonic and magmatic evolution of the Andes of do Paraná, Reavaliação da potencialidade e prospectivi-
northern Argentina and Chile: Earth Science Reviews, v. dade em hidrocarbonetos: Pauli Petro Consorcio, Sao
18, p. 303–332. Paulo, p. 19–40.
Cole, G. A., R. J. Drozd, R. A. Sedivy, and H. I. Halpern, 1987, Gleadow, A. J. W., I. R. Duddy, and J. F. Lovering, 1983,
Organic geochemistry and oil–source correlations, Fission track analysis: a new tool for the evaluation of
Paleozoic of Ohio: AAPG Bulletin, v. 71, p. 788–809. thermal histories and hydrocarbon potential: American
Copper, P., 1977, Paleolatitudes in the Devonian of Brazil and Petroleum Engineers Association Journal, v. 23, p. 93–102.
the Frasnian-Famennian mass extinction: Palaeogeog- Goulart, E. P., and Jardim, N. S., 1982, Avaliação geoquímica
raphy, Palaeoclimatology, Palaeoecology, v. 21, das formações Ponta Grossa e Irati-bacia do Paraná, in R.
p. 165–207. Yoshida and E. Gamma, Jr., eds., Geologia da bacia do
Dalmayrac, B., 1988, Characteres generales de la evolucion Paraná, Reavaliação da potencialidade e prospectividade
geologica de los Andes Peruanos: Instituto Geologica em hidrocarbonetos: Pauli Petro Consorcio, Sao Paulo,
Minero y Metalurgica, Buletin 12, Series D Estudios Espe- p. 41–74.
ciales, Lima, Peru. Grahn, Y., and M. V. Caputo, 1992, Early Silurian glaciations
Dallmus, K. F., 1958, Mechanics of basin evolution and its in Brazil: Palaeography, Palaeoclimatology, Palaeoe-
relation to the habitat of oil in the basin, in L. G. Weeks, cology, v. 99, p. 9–15.
ed., Habitat of oil: AAPG, p. 883–931. Hatcher, R. D., 1988, The third synthesis: Wenlock to mid-
Dewey, J. F., 1988, Extensional collapse of orogens: Tectonics, Devonian (end of Acadian orogeny), in A. L. Harris and
v. 7, p. 1123–1139. D. J. Fettes, eds., The Caledonian-Appalachian orogen:
DeWit, M. J., and I. G. D. Ransome, 1992, Regional inversion Geological Society of London Special Publication 38,
tectonics along the southern margin of Gondwana, in M. J. p. 499–506.
DeWit and I. G. D. Ransome, eds., Inverson tectonics of Helwig, J., 1972, Stratigraphy, sedimentation, paleogeog-
the Cape fold belt, Karoo and Cretaceous basins of raphy, and paleoclimates of Carboniferous (“Gondwana”)
southern Africa: Rotterdam, Balkema, p. 15–21. and Permian of Bolivia: AAPG Bulletin, v. 56,
Dingle, R. V., W. G. Siesser, and A. R. Newton, 1983, p. 1008–1033.
Mesozoic and Tertiary geology of southern Africa: Herve, F., E. Godoy, M. A. Parada, V. A. Ramos, C. Rapela, C.
Rotterdam, Balkema, 375 p. Mpodozis, and J. Davidson, 1987, A general view of the
Elliot, D. H., 1975, Gondwana basins of Antarctica, in K. S. W. Chilean-Argentine Andes, with emphasis on their early
Campbell, ed., Third Gondwana symposium: Canberra, history, in J. Monger and J. Francheteau, eds., Circum
Australia, p. 493–536. Pacific orogenic belts and evolution of the Pacific Ocean
98 Williams

basin: American Geophysical Union Geodynamics Series, Marshak, S., 1986, Structure and tectonics of the Hudson
v. 18, p. 97–113. valley fold-thrust belt, eastern New York: GSA Bulletin,
Illich, H. A., F. R. Haney, and M. Mendoza, 1981, Geochem- v. 97, p. 354–368.
istry of oil from Santa Cruz basin, Bolivia: case study of Martinez, C., 1980, Geologie des Andes Boliviennes: Travaux
migration-fractionation: AAPG Bulletin, v. 65, et Documents de L’Orstom, Bondy, France, 353 p.
p. 2388–2402. Matchus, E. J., and T. S. Jones, 1984, Co-Chairmen of the
Isaacson, P. E., 1975, Evidence for a western extracontinental Stratigraphic Committee of the West Texas Geological
land source during the Devonian Period in the central Society, East-west cross-section through the Permian basin
Andes: GSA Bulletin, v. 86, p. 39–46. of West Texas: West Texas Geological Society Publication
Karner, G. D., and J. F. Dewey, 1986, Rifting: lithospheric 84-79, three plates.
versus crustal extension as applied to the Ridge basin of Mayer, L., 1987, Subsidence analysis of the Los Angeles
southern California, in M. T. Halbouty, ed., Future basin, in R. V. Ingersoll and W. G. Ernst, eds., Cenozoic
petroleum provinces of the world: AAPG Memoir 40, basin development of coastal California: Proceedings of
p. 317–337. the Rubey Colloquim Series, Vol. 6, Englewood Cliffs,
King, P. B., 1977, The Evolution of North America, Revised New Jersey, Prentice-Hall, p. 300–320.
Edition: Princeton University Press, Princeton, New McKenzie, D., 1978, Some remarks on the development of
Jersey, 197 p. sedimentary basins: Earth and Planetary Science Letters,
Klein, G. DeV., and J. B. Kupperman, 1992, Pennsylvanian v. 40, p. 25–32.
cyclothems: methods of distinguishing tectonically Messner, J. C., and L. C. P. Wooldridge, 1964, Maranhao
induced changes in sea level from climatically induced Paleozoic basin and Cretaceous coastal basins, northern
changes: GSA Bulletin, v. 104, p. 166–175. Brazil: AAPG Bulletin, v. 48, p. 1475–1512.
Kleinschmidt, G., and F. Tessensohn, 1987, Early Paleozoic Milani, E. J., 1992, Intraplate tectonics and the evolution of
westward-directed subduction at the Pacific margin of the Paraná basin, S.E. Brazil, in M. J. DeWit and I. G. D.
Antarctica, in G. D. McKenzie, ed., Gondwana 6: structure, Ransome, eds., Inversion tectonics of the Cape fold belt,
tectonics, and geophysics: American Geophysical Union, Karoo and Cretaceous basins of South Africa: Rotterdam,
Geophysical Monograph 40, p. 89–105. Balkema, p. 101–108.
Lambeck, K., S. Cloetingh, and H. McQueen, 1987, Intraplate Miller, H., 1981, Pre-Andean orogenies of southern South
stresses and apparent changes in sea level: the basins of America in the context of Gondwana, in M. M. Cresswell
northwestern Europe, in C. Beaumont and A. J. Tankard, and P. Vella, eds., Gondwana Five: Proceedings of the
eds., Sedimentary basins and basin-forming mechanisms: Fifth International Gondwana symposium: Rotterdam,
Canadian Society of Petroleum Geologists Memoir 12, Balkema, p. 237–242.
p. 259–268. Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1972,
Lash, G. G., 1987, Geodynamic evolution of the lower Sierras subandinas, in J. C. M. Turner, ed., Geologia
Paleozoic central Appalachian foreland basin, in: C. regional Argentina: Cordoba, Academia Nacional de
Beaumont and A. J. Tankard, eds., Sedimentary basins Ciencas I, p. 95–138.
and basin-forming mechanisms: Canadian Society of Mosmann, R., F. U. H. Falkenhein, A. Goncalves, and F. N.
Petroleum Geologists Memoir 12, p. 413–423. Filho, 1986, Oil and gas potential of the Amazon Paleozoic
Lawver, L. A., and C. R. Scotese, 1987, A revised reconstruc- basins, in M. T. Halbouty, ed., Future Petroleum Provinces
tion of Gondwanaland, in G. D. McKenzie, ed., Gondwana of the World: AAPG Memoir 40, p. 207–241.
six: structure, tectonics, and geophysics: American Northfleet, A. A., R. A. Medeiros, and H. Muhlmann, 1969,
Geophysical Union, Geophysical Monograph 40, p. 17–22. Reavaliação dos dados geologicos da bacia do Paraná:
Lerche, I., 1990, Basin Analysis: Quantitative Methods, Vol. 1: Bolletin Tecnico PetroBras, Rio de Janeiro, v. 12,
San Diego, Academic Press, 562 p. p. 291–346.
Levell, B. K., J. H. Braakman, and K. W. Rutten, 1988, Oil- Padula, E. L., 1972, Subsuelo de la Mesopotamia y regiones
bearing sediments of Gondwana glaciation in Oman: adyacentes, in J. C. M. Turner, ed., Geologia Regional de
AAPG Bulletin, v. 72, p. 775–796. Argentina, Cordoba, Academia Nacional de Ciencias,
p. 213–235.
Lohmann, H. H., 1970, Outline of tectonic history of Bolivian
Padula, V. T., 1969, Oil shale of Permian Irati Formation,
Andes: AAPG Bulletin, v. 54, p. 735–757.
Brazil: AAPG Bulletin, v. 53, p. 591–602.
Lopez-Gamundi, O. R., 1989, Postglacial transgressions in
Pezzi, E. E., and M. E. Mozetic, 1989, Cuencas sedimentarias
late Paleozoic basins of western Argentina: a record of
de la region Chacoparanense, in G. A. Chebli and L. A.
glacioeustatic sea level rise: Palaeogeography, Palaeocli-
Spalletti, eds., Cuencas sedimentarias Argentinas: Serie
matology, Palaeoecology, v. 71, p. 257–270.
correlacion geologia 6, Universidad Nacional de
Lopez-Gamundi, O. R., and E. A. Rossello, 1993,
Tucuman, Instituto Superior de Correlacion Geologia,
Devonian–Carboniferous unconformity in Argentina and
p. 65–78
its relation to the Eo-Hercynian orogeny in southern
Quinlan, G. M., 1987, Models of subsidence mechanisms in
South America: Geologische Rundschau, v. 82, p. 136–147.
intracratonic basins and their applicability to North
Lucazeau, F., and S. LeDouaran, 1984, Numerical model of American examples, in C. Beaumont and A. J. Tankard,
sediment thermal history, comparison between the Gulf of eds., Sedimentary basins and basin-forming mechanisms:
Lion and the Viking Graben, in B. Durand, ed., Thermal Canadian Society of Petroleum Geologists Memoir 12,
phenomena in sedimentary basins: Paris, Editions p. 463–481.
Technip, p. 211–217. Quinlan, G. M., and C. Beaumont, 1984, Appalachian
Malizia, D. C., E. E. Miller, M. Del R. Rosso, I. Labayen, and thrusting, lithospheric flexure, and the Paleozoic stratig-
S. Hernaez, 1993, Geochemical interpretation of the raphy of the eastern interior of North America: Canadian
Carboniferous–Permian in the alhuampa sub-basin, Journal of Earth Sciences, v. 21, p. 973–996.
northeast Argentina: Comptes Rendus XII ICC-P, v. 1, Ramos, V. A., 1986, Discussion, tectonostratigraphy, as
p. 239–252. applied to analysis of South American Phanerozoic basins:
Tectonic Subsidence Analysis and Paleozoic Paleogeography of Gondwana 99

Transactions of the Geologic Society of South Africa, v. 89, D. McKenzie, ed., Gondwana six: structure, tectonics, and
p. 427–429. geophysics: American Geophysical Union, Geophysical
Ramos, V. A., 1989, The birth of southern South America: Monograph 40, p. 271–286.
American Scientist, v. 77, p. 444–450. Urien, C. M., J. J Zambrano, and L. R. Martins, 1981, The
Ramos, V. A., T. E. Jordan, R. W. Allmendinger, C. basins of southeastern South America (southern Brazil,
Mpodozis, S. M. Kay, J. M. Cortes, and M. Palma, 1986, Uruguay and eastern Argentina) including the Malvinas
Paleozoic terranes of the central Argentine-Chile Andes: plateau and southern South Atlantic, paleogeographic
Tectonics, v. 5, p. 855–880. evolution: Comite Suderamerico del Jurasico y Cretacico:
Ronov, A., V. Khain, and K. Seslavinsky, 1984, Atlas of litho- Cuencas sedimentarias del Jurasico y Cretacico del Sur,
logical and paleogeographical maps of the World, Late v. 1, p. 45–126.
Precambrian and Paleozoic of continents: Leningrad, Van Houten, F. B., and R. B. Hargraves, 1987, Palaeozoic drift
U.S.S.R. Academy of Sciences, 70 p. of Gondwana: palaeomagnetic and stratigraphic
Schmoker, J. W., 1986, Oil generation in the Anadarko basin, constraints: Geological Journal, v. 22, p. 341–359.
Oklahoma and Texas: modeling using Lopatin’s method: Van Der Voo, R., 1988, Paleozoic paleogeography of North
Oklahoma Geological Survey Special Publication 86-3, America, Gondwana, and intervening displaced terranes:
40 p. comparisons of paleomagnetism with paleoclimatology
Sclater, J. C., and P. A. F. Christie, 1980, Continental and biogeographical patterns: GSA Bulletin, v. 100,
stretching: an explanation of the post-mid-Cretaceous p. 311–324.
subsidence of the North Sea basin: Journal of Geophysical Vargas, J. M., 1988, Petroleum potential of the Huallaga
Research, v. 85, p. 3711–3739. basin, eastern Peru: III Simposio Bolivariano, p. 196–225.
Scotese, C. R., R. K. Bambach, C. Barton, R. Van Der Roo, and Veevers, J. J., 1980, Basins of the Australian craton and
A. M. Ziegler,1979, Paleozoic base maps: Journal of margin, in A. W. Bally, P. L. Bender, T. R. McGetchin, and
Geology, v. 87, p. 217–277. R. I. Walcott, eds., Dynamics of plate interiors: geody-
Sempere, T., 1995, Phanerozoic evolution of Bolivia and namics series: American Geophysical Union and Geolog-
adjacent regions, in A. J. Tankard, R. Suarez, and H. J. ical Society of America, v. 1, p. 73–80.
Welsink, Petroleum basins of South America: AAPG Veevers J. J., and C. McA. Powell, 1987, Late Paleozoic glacial
Memoir 62, this volume. episodes in Gondwanaland reflected in transgressive-
Soares, P. C., P. M. B. Landim, O. Sinelli, E. Wernick, F. T. regressive depositional sequences in Euramerica: GSA
Wu, and A. P. Fiori, 1977, Associações litológicas do Bulletin, v. 98, p. 475–487.
subgrupo Itararé e sua interpretação ambiental: Revista Veevers, J. J., C. McA. Powell, J. W. Collinson, and O. R.
Brasileira de Geosciencias, v. 7, p. 131–149. Lopez-Gamundi, in press, Synthesis, in J. J. Veevers and
Starck, D., 1995, Silurian–Jurassic stratigraphy and basin C. McA. Powell, eds., Permian–Triassic Pangean basins
evolution of northwestern Argentina, in A. J. Tankard, R. and foldbelts along the Panthalassian margin of
Suarez, and H. J. Welsink, Petroleum basins of South Gondwana: GSA Memoir 184.
America: AAPG Memoir 62, this volume. Visser, J. N. J., 1987, The influence of topography on the
St. John, B., 1986, Antarctica—geology and hydrocarbon Permo-Carboniferous glaciation in the Karoo basin and
potential, in M. T. Halbouty, ed., Future petroleum adjoning areas, South Africa, in G. D. McKenzie, ed.,
provinces of the world: AAPG Memoir 40, p. 56–100. Gondwana six: stratigraphy, sedimentology, and paleon-
Stump, E., 1987, Construction of the Pacific margin of tology: American Geophysical Union, Geophysical
Gondwana during the Pannotois cycle, in G. D. McKenzie, Monograph 40, p. 123–129.
ed., Gondwana six: structure, tectonics, and geophysics: Visser, J. N. J., 1993, The tectono-geographic evolution of part
American Geophysical Union, Geophysical Monograph of southwestern Gondwana during the Carboniferous and
40, p. 77–87. Permian: Comptes Rendus XII ICC-P, v. 1, p. 447–454.
Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M. Voorhoeve, H., and G. Houseman, 1988, The thermal
Turic, A. B. Franca, E. J. Milani, B. B. de Brito Neves, N. evolution of lithosphere extending on a low-angle detach-
Eyles, H. de Santa Ana et al., 1995, Tectonic controls of ment zone: Basin Research, v. 1, p. 1–9.
basin evolution in southwestern Gondwana, in A. J. Waples, D. W., H. Kamata, and M. Suizu, 1992, The art of
Tankard, R. Suarez, and H. J. Welsink, Petroleum basins maturity modeling, part 1: finding a satisfactory geologic
of South America: AAPG Memoir 62, this volume. model: AAPG Bulletin, v. 76, p. 31–46.
Theron, J. N., and J. C. Loock, 1988, Devonian deltas of the Watts, A. B., 1981, The U.S. Atlantic continental margin,
Cape Supergroup, South Africa, in N. J. McMillan, A. F. subsidence history, crustal structure, and thermal
Embry, and D. J. Glass, eds., Devonian of the world: evolution: AAPG Short Course Notes, 85 p.
Canadian Society of Petroleum Geologists Memoir 14, v. Watts, A. B., and J. Thorne, 1984, Tectonics, global changes in
1, p. 729–739. sea level, and their relationship to stratigraphic sequences
Thomas, W. A., and P. E. Schenk, 1988, Late Paleozoic sedi- at the U.S. Atlantic continental margin: Marine and
mentation along the Appalachian orogen, in A. L. Harris Petroleum Geology, v. 1, p. 319–338.
and D. J. Fettes, eds., The Caledonian-Appalachian Watts, A. B., and M. S. Steckler, 1981, Subsidence and
orogen: Geological Society of London Special Publication tectonics of Atlantic-type continental margins: Oceologica
38, p. 515–530. Acta 1981, p. 143–153.
Thorne, J. A., and A. Watts, 1989, Qualitative analysis of Wilner, A. P., U. S. Lottner, and H. Miller, 1987, Early
North Sea subsidence: AAPG Bulletin, v. 73, p. 88–116. Paleozoic structure development in the N.W. Argentine
Tissot, B. P., R. Pelet, and P. Ungerer, 1988, Thermal history basement of the Andes and its implication for geodynamic
of sedimentary basins, maturation indices, and kinetics of reconstructions, in G. D. McKenzie, ed., Gondwana six:
oil and gas generation: AAPG Bulletin, v. 71, p. 1445–1466. structure, tectonics, and geophysics: American Geophys-
Uliana, M. A., and K. T. Biddle, 1987, Permian to Late ical Union, Geophysical Monograph 40, p. 229–238.
Cenozoic evolution of northern Patagonia: main tectonic Wilson, K. M., M. J. Rosol, W. W. Hay, and G. A. Harrison-
events, magmatic activity, and depositional trends, in G. Christopher, 1989, New model for the tectonic history of
100 Williams

west Antarctica: a reappraisal of the fit of Antarctica in Author’s Mailing Address


Gondwana: Eclogae Geologicae Helvetiae, v. 82, p. 1–28.
Wopfner, H., 1981, Development of Permian intracratonic Kenneth E. Williams
basins in Australia, in M. M. Cresswell and P. Vella, eds., Texaco Inc.
Gondwana Five: Proceedings of the Fifth International 3901 Briarpark
Gondwana Symposium: Rotterdam, Balkema, p. 185–190. Houston, Texas 77402
Yacimentos Petroliferos Fiscales Bolivianos, 1972, Resumen U.S.A.
de la Geologia Petrolera de Bolivia: YPFB, 92 p.
Zalan, P. V., S. Wolff, M. A. M. Astolfi, I. S. Vieira, J. C. J.
Concelcao, V. T. Appi, E. V. S. Neto, J. R. Cerqueira, and
A. Marques, 1990, The Paraná basin, Brazil in Leighton, M.
W., D. R. Kolata, D. F. Oltz and J. J. Eidel, eds., Interior
cratonic basins: AAPG Memoir 51, p. 681–708.
Mesozoic–Cenozoic Andean Paleogeography and
Regional Controls on Hydrocarbon Systems

James L. Pindell Kenneth D. Tabbutt


Department of Earth Sciences Evergreen State College
Dartmouth College Olympia, Washington, U.S.A.
Hanover, New Hampshire, U.S.A.

Abstract

P alinspastic paleogeographic maps of western and northern South America, including the entire 8500-km
“Andean system” from Trinidad to Cape Horn, are presented for nine Mesozoic–Cenozoic time
intervals. The maps show (1) the spatial record of formational lithostratigraphic units; (2) continental, shallow
marine, and deeper marine paleoenvironments and the location of active magmatic arcs through time; (3)
progressive structural and tectonic development; (4) relative motions of adjacent plates affecting the Andes;
and (5) paleolatitude. Phases and causes of geologic development are summarized from the maps and other
information. Depositional systems are related to tectonic evolution, with implications drawn for hydrocarbon
systems and history. It is shown that tectonic, depositional, and hydrocarbon histories are closely interrelated,
having occurred in fairly discrete pulses through time, each with its own significance to hydrocarbon potential.

Resumen

S e presentan mapas paleogeográficos palinspasticos de Suramerica septentrional y occidental, de todo el


“Sistema Andino” de 8500 km de largo desde Trinidad hasta el Cabo de Horno, para 9 intervalos de
tiempo del Mesozoico-Cenozoico. Los mapas muestran: La distribución de las unidades litostratigráficas
formacionales; paleoambientes continentales, marinos someros, y marinos mas profundos y la ubicación de los
arcos magmáticos a través del tiempo; el desarrollo progresivo estructural y tectónico; los movimientos
relativos de las placas adyacentes afectando los Andes; y las paleolatitud. Se resumen las fases y las causas del
desarrollo geológico a través de los mapas y de otra información. Se relacionan los sistemas deposicionales a la
evolución tectonica, y sus implicaciones para los sistemas y la historia hidrocarburifera. Las historias tectónicas,
deposicionales, y hidrocarburiferas estan intimamente relacionadas, habiendo occurrido en pulsos bastante-
mente discretos a través del tiempo, cada uno con su propio significado para el potencial hidrocarburifero.

INTRODUCTION tion of regional and local Phanerozoic geologic histories


and integration with modern concepts of geologic
During the period 1988–1992, we conducted an processes, (5) in-depth assessments of Neogene–Recent
extensive and comprehensive analysis of the Andean “Andean phase” strain and orogenic history, and (6)
system of western and northern South America (Pindell considerations of hydrocarbon history and habitat along
et al., 1992) to refine the paleogeographic evolution, to strike. The present paper presents an abbreviated version
define the main tectonic and sedimentary relationships of the aforementioned study’s Mesozoic–Cenozoic paleo-
controlling basin development, and to assess the hydro- geographic assessment, regional tectonic history, and
carbon history and habitat along the chain. That study aspects of basin development and hydrocarbon history.
included the following elements: (1) up-to-date assess- Much of the length of the Andes has been the site of
ments of relative and absolute plate motions, (2) creation arc magmatism during Mesozoic and Cenozoic time. It is
of an integrated paleogeographic database and plotting therefore useful to think of the Andes as an evolving arc
program with over 8400 lithochronologic control points system whose development was dominated by arc
compiled mainly from the public domain and our processes which were in turn controlled by the motion of
personal studies, (3) palinspastic reconstruction of South America relative to the mantle and the adjacent
Phanerozoic compressional and extensional bulk strain plates. We find that, in general, the concepts on arc
and block displacements along the chain, (4) documenta- behavior synthesized by Dewey (1980) describe quite

Pindell, J. L., and Tabbutt, K. D., 1995, Mesozoic–Cenozoic Andean paleogeography and 101
regional controls on hydrocarbon systems, in A. J. Tankard, R. Suárez S., and H. J.
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 101–128.
102 Pindell and Tabbutt

well the successive periods of historical development of 2. The Amaime-Chaucha complex of the western
the Andean chain. Dewey’s arc models stress the signifi- flank of the central Cordillera of Colombia and
cance of a kinematic framework defined by the position northern Ecuador, Early Cretaceous accretion
of the subducting slab in the mantle. The motion of the (Aspden and McCourt, 1986) or Late Cretaceous
overriding plate (South America) relative to the slab accretion, preferred here (see Pindell and Erikson,
helps to control many elements of the arc, such as the dip 1994) (Figures 5, 6)
of the Benioff zone, topographic expression of the arc 3. The western Cordillera of Colombia, latest Creta-
system, convergent or divergent tectonics within the arc, ceous–early Paleogene accretion (Bourgois et al.,
back-arc extension or foreland thrusting, chemistry and 1987; Daly, 1989; Pindell and Erikson, 1994)
intensity of magmatism, erodability of the arc as a source (Figure 6)
of sediment, and sense of shear at trench-parallel faults 4. The Panamanian arc and basement (Choco
within the arc. Synthesis of these types of geologic obser- terrane), Eocene?–Recent accretion (Pindell and
vations at arc segments over time can be used to infer Dewey, 1982; Grosser, 1989; Pindell, 1993)
past aspects of subduction and to iteratively deduce pale- (Figures 7–9)
ogeographic development. These in turn can be applied 5. The Pinón Formation and Costa Province, Late
to understanding plate boundary history and basinal Cretaceous–early Paleogene accretion (Daly, 1989)
dynamics, with direct implications for hydrocarbon (Figures 3, 6)
history and habitat. The synopsis presented here adheres 6. Macuchi arc, northern Ecuador (Daly, 1989)
to this philosophy. It is useful for understanding regional (Figure 5)
Andean evolution, arc behavior over time, and Andean 7. Rocas Verdes complex, southern Chile, Late Creta-
basinal dynamics with implications for hydrocarbon ceous back-arc closure and arc accretion (Dalziel,
systems. 1986) (Figures 2–5).

Prior to their respective times of accretion, these terranes


PALEOGEOGRAPHIC MAPS were not a part of the autochthonous South American
shield and are shown on the paleogeographic maps as
The goal of creating paleogeographic maps (Figures arriving or developing at the various times due to
1–9) is to portray the relative positions of plates, paleo- relative plate motions.
sedimentation patterns and environments, structural and To assess extension and shortening, we broke the
tectonic features, paleolatitude, and other information as Andes south of central Colombia into 23 segments along
it existed in the past. To improve the accuracy of spatial strike. We then assigned estimated azimuths and values
reconstruction through time, it is necessary to estimate of shortening or extension for each segment by tectonic
and restore bulk strain by working backward in time so event. Strain values were estimated for each segment
that the maps represent palinspastic reconstructions of working back in time so that values were compounded
geologic development. Creation of such maps involves a for each previous tectonic event. For example, values for
fairly straightforward process of integration (Ziegler et the late Miocene–Recent shortening range from 20–35
al., 1985; Pindell, 1985). Nevertheless, the methodology km for segments in Ecuador and southern Patagonia to
we used to estimate bulk strain in the palinspastic recon- 250–300 km for segments in the central Andes. Older
structions deserves some explanation. shortenings were then added to these values for the
appropriate times. Shortening was approximately
Palinspastic Reconstruction removed on the maps by homogeneously stretching each
segment in the direction opposite to shortening, as
Satisfactory assessment of former terrane accretion, estimated by structure and plate motions, if known.
back-arc and inter-arc extension, intra-arc and foreland Likewise, judging from red bed thicknesses and subse-
shortening, and margin-parallel slip are the main quent sedimentation rates due to thermal subsidence,
elements of bulk palinspastic reconstruction of strain. For Triassic–Jurassic extension of a few tens of kilometers is
the Andes, our objective is to recreate the original estimated for Ecuador to northern Argentina, ranging to
geographic and geometric attributes of depocenters and about 100 km for both the central Colombian segment
structural provinces through time. This allows geologic (eastern Cordilleran back-arc basin; Figures 2, 3) and the
processes and systems to be more readily inferred collective rifts of southern Patagonia (Figures 1–3). In
relative to inferences derived from the present-day these cases, extension was removed by closing individual
deformed geologic record. rifts between transfer zones.
Accreted terranes and episodes of Mesozoic–Cenozoic Margin-parallel and other strike-slip offsets must also
accretion include the following: be nested into the successive palinspastic restorations.
For example, the pre-Neogene reconstruction of the
1. The Caribbean nappes of northern South America, numerous blocks between high-strain zones in the
diachronous Maastrichtian–Miocene accretion northern Andes (central Colombia to western Venezuela)
(MacDonald et al., 1971; Tschanz et al., 1974; is achieved by finite-difference solution (Dewey and
Stephan et al., 1980; Case et al., 1984; Avé- Pindell, 1985, 1986). The maps restore 110 km of dextral
Lallemant, 1990; Pindell and Barrett, 1990) shear and 25 km of shortening in the Merida Andes, 115
(Figures 7–9) km of sinistral shear along the Santa Marta fault, 30 km
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 103

of total shortening across Sierra Perijá, and 65 km of Next, the base map geographic features, the struc-
dextral shear along the Oca fault, west of the Perijá tures, and the lithologic data were palinspastically
(Figure 9). Finite-difference solution with these values restored as previously outlined. The differences in the
yields a total shortening in the northernmost eastern positions of the faint and bolder geographic lines allow
Cordillera of Colombia in excess of 150 km, with a net visual observation of the palinspastic differences for each
direction toward the east-southeast. segment along strike. Structural lines crossing segment
In a system with as many diverse opinions on boundaries with different shortening values were then
evolution as the Andes, particularly in light of the made continuous by smoothing any offset that might
multiple deformation events, it is unrealistic to expect have existed after restoration. Paleoenvironment bound-
different methods or biases of estimating strain to yield aries were drawn separating regions of equivalent paleo-
bulk strain values through time that agree with high environment, indicated by the patterns, as defined by the
precision. We can only strive at this point for internal raw lithochronologic data. Paleolatitudes (after Scotese et
consistency in the methods applied, ensuring that the al., 1987), key cross sections, notes on paleogeographic
values used are in line with most others offered. This development, tectonic processes, and sandstone prove-
study has integrated numerous types of information in nance were then added to complete the 1:5,000,000 scale
the hopes of satisfying the majority of data. We are thus maps. The Mesozoic and Cenozoic intervals of these
confident that the successive palinspastic restorations maps were then reduced and simplified to create Figures
provide an autochthonous framework of features for the 1–9 in this paper. Lithochronologic data, notes, and cross
past that is far closer to former reality than the present- sections were removed, and formation names of regional
day geography of features and that allows a much more extent were added, with notations for source and
accurate depiction of the original shapes and former rela- reservoir units. Thus, the maps in Figures 1–9 are graphi-
tionships of basins and uplifts. Successive palinspastic cally as accurate as the larger versions but lack the
restorations can be seen on the maps by shifts in the control points, lithologic designations, and interpreta-
positions of the faint (present-day) and bolder (former) tional detail.
geographic lines.
Phases of Paleogeographic Evolution
Construction of the Maps
Primary Andean developments, their causal mecha-
Figures 1–9 are simplified reductions of more detailed nisms, and their impact on Andean hydrocarbon systems
maps created in Mercator projection at 1:5,000,000. On are summarized in Table 1 and are more fully outlined in
palinspastically restored base maps of western and the following paragraphs.
northern South America, these simplified maps show the Beginning in the Late Permian in the central Andes
following features: (1) the migration of plates and and continuing along the entire Andes by the Triassic,
terranes that have affected the Andes, (2) paleosedimen- back-arc extension formed a series of restricted troughs
tation patterns by paleoenvironment and formation (extensional arc in the sense of Dewey, 1980) in which red
name, (3) designation of main source rock and reservoir beds, arc volcanics, and marine strata accumulated
rock intervals, (4) primary structural and tectonic (Figure 1). Some of the marine units such as the Santiago
features, (5) the trace of arc magmatism, and (6) paleolati- and Pucara are, or were, oil-prone source rocks whose
tude. Faint geographic lines (coast lines, political bound- deposition probably fits models of rifted basins with
aries, and lakes) are present-day positions; bolder equiv- marine access and restricted circulation. Other rift basins
alents are relative paleopositions of the same. Depicted farther south did not experience a marine transgression
paleoenvironments include continental, shallow marine and instead accumulated nonmarine organic-rich source
(less than ~200 m paleowater depth), deeper marine rock units such as the Los Molles Formation of the
(greater than ~200 m), and the axis of the magmatic arc, Neuquén basin and, a little later, the Pozo D-129 unit of
which takes preference over other paleoenvironments the San Jorge basin (Figures 2, 3). In the north, westward-
(e.g., the shallow marine Triassic Payande unit of Colom- propagating intracontinental rifting between Yucatan
bia is covered by the arc pattern) (Figure 1). and Venezuela heralded the Proto-Caribbean seaway,
The maps were created as follows. Phanerozoic time which opened in a fan-like fashion (Figures 1, 2). In
(including the Paleozoic) was divided into 16 intervals Patagonia, a series of en-echelon rifts developed along
for which a general tectonic or depositional style preexisting structural trends associated with pre-
prevailed. The age of each map (in Ma) is shown in its Mesozoic accretion of arc terranes (Figures 1–3). This
key and falls about midway within the time increment. period of extension and breakup is directly tied to the
On geographic base maps, all of the lithochronologic breakup of Pangea: Andean back-arc extension was
data in the databases (lithology, thickness, formation, probably due to subduction zone rollback, with which
and contact relationships) within a given map’s time the South American craton did not keep pace (Dewey,
increment, both isotopic and stratigraphic, were plotted 1980). All of the back-arc troughs along the Andes
in present-day coordinates. Next, structures with known became sites of thermal subsidence in the Middle
ages matching the various time increments were added Jurassic–Early Cretaceous (Figure 2), depending on
in present-day coordinates. At this point, the maps were location. Arc morphology was subdued, typically with
present-day plots of lithologic and paleostructural data volcanic islands surrounded variably by marine
for each increment in time. seaways.
104 Pindell and Tabbutt

Table 1—Main Andean Developments, Causes, and Impact on Hydrocarbon Systems

Andean Driving Cause of Impact on Hydrocarbon


Development Development Systems
Triassic–Jurassic Subduction zone rollback, Deposition of nonmarine/
rift basins along arc axis rifting from Yucatan marine source rocks in rift
and northern passive margin. during Pangean breakup. basins from Ecuador south.

Early Cretaceous expansion Thermal subsidence after Transgressive reservoir sand-


of epicontinental seas and area rifting, rising long-term stone deposition, carbonate
of sedimentation. eustasy. deposition, locally reservoir.

Albian onset of convergent Aptian onset of north Africa– Initial (Albian) source rock
tectonism and foredeep geometry South America separation deposition,central Andean
in central Andes. in equatorial Atlantic, foreland, continue transgressive
initiate convergent arc. reservoirs.

Middle Cretaceous drowning High long-term sea level, Best regional source rock int-
of northern shelf platform, distant strand lines, erval deposited regionally, in
expansion and deepening of starvation of clastic areas intersected by oxygen
epicontinental seas elsewhere. materials from depocenters. minimum zone.

Late Cretaceous arc collision Westward drift of South End source rock deposition,
and onset of east-dipping America closed back-arcs, begin progradational reservoir
subduction in north and south, convergence continued by rock deposition as seas
begin foredeep geometry in Col- subduction at new Benioff regressed. Possible local onset
ombia/western Venezuela and zones. Late Cret lowering of maturation at thrust zones
southern Chile/Argentina, infilling of long-term sea level. and sedimentary buildups.
of epicontinental seaways.

Paleogene removal of seas from Incaic uplift due to W accel- Deposition of important
much of craton, uplift and creation eration of South America fluvial reservoirs in foreland
of continuous Andean barrier, begin across mantle, rapid sub- areas (sub-Andes), major
large scale eastward shedding of duction rates. Onset of phase of maturation and east-
detritus from Andes. Begin eastward oblique Caribbean collision. ward oil migration. Eastward
migrating foredeep in Venezuela. migrating maturation
mechanism in Venezuela.

Late middle Cenozoic waning of Slowing of South America Filling of Incaic structures by
orogenesis and foreland develop- across the mantle and of migration from Incaic kitchens,
ment. Continue Caribbean fore- subduction. Continue deposition of additional fluvial
deep migration. Caribbean–South America reservoirs in foreland.
relative motion.

Late Cenozoic rejuvenation Acceleration of South Second primary maturation


of orogenesis (Andean), rapid America across mantle phase in Andes, eastward
uplift, molassic deposition and and of subduction rate, migration of thrust front/foredeep
foredeep development. Carib- and, in north, intensification axis and local eastward jumping
bean foredeep now in Maturin of Panamanian collision. of foreland deformation.
and Trinidad area. Progressive younging of Primary maturation phase in
crust entering Andes trench. eastern Venezuela Basin.

For much of the Andes, epicontinental marine and oceanic crust (De Wit, 1977). Such extension appears to
nonmarine deposition east of the arc dominated the have been continuous from the Jurassic, still related to
Early Cretaceous, indicating generally stable tectonic rollback. In the north, southward transgression across the
conditions typified by mainly thermal subsidence passive portion of the margin was underway. Probable
(Figure 3). However, extensional tectonism persisted southward-propagating opening of a back-arc basin
along the Chilean portion of the back-arc, with the Rocas along Colombia left the central Cordillera as a remnant
Verdes basin of southern Chile developing a basement of arc passive margin after Triassic–Middle Jurassic
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 105

plutonism had ceased (Figures 3, 4). The Andean arc, preservation of organic-rich shales (e.g., Napo and
which was ensialic in the central Andes, probably trended Chonta formations). Farther south, in Bolivia and
offshore at Ecuador to connect with the Amaime- Argentina, tectonism and associated lithospheric flexure
Chaucha arc terrane (Figures 3, 4). These continued north were relatively less and shallow marine to nonmarine
into the Greater Antilles arc and the North American arcs molasse mainly accumulated.
of Mexico and the Chortis block to the northwest By the end of the Cretaceous (Figure 6), falling long-
(Pindell, 1993). term sea level and infilling of epicontinental accommo-
During the Aptian, this Aleutian-like intraoceanic arc dation space heralded the removal of the seas from the
apparently flipped polarity, with the associated creation craton, with generally regressive deposits marking the
of blueschists and other metamorphic rocks at the onset migrating marine margins. The Campanian(?) accretion
of west-dipping Aptian–Albian subduction along the of the Amaime-Chaucha terrane (Pindell and Erikson, in
eastern side of the arc. These metamorphic rocks are now press) loaded the central Cordillera and created the
found in obducted masses all around the Caribbean and Umir-Colon basin to the east with asymmetric foredeep
along the eastern limit of the Amaime terrane in geometry. The deposits in this basin prograded north-
Colombia. From the Albian, the arc migrated eastward ward toward the Proto-Caribbean seaway, leaving a
toward the remnant Colombian and northern Ecuador- paralic to nonmarine depocenter in the wake of the
ian margin and the Proto-Caribbean seaway to the north. marine basin. After accretion of the Amaime, continued
Plutonism in this east-facing arc is well known in the South American–Caribbean plate convergence in the
Antilles. Likewise, the Buga batholith of the Amaime northern Andes was accommodated by initiation of east-
terrane (~99 Ma) (Aspden and McCourt, 1986) may dipping subduction of Caribbean crust outboard of the
represent Antilles equivalent arc magmatism during terrane. Progressive accretion of offscraped upper
closure of the Colombian back-arc (Pindell, 1993; Pindell Caribbean plate materials probably formed most of
and Erikson, in press). Colombia’s western Cordillera, west of the already
By the end of Albian time (Figure 4), marine water accreted Amaime terrane. The eastward continuation of
depths in the north were sufficient for initiation of source the Panamanian arc (Figure 6) defined the western and
rock deposition and preservation. South America was southern margins of the Caribbean plate from the
migrating toward the Amaime terrane. Peruvian orogen- Santonian or the Campanian (Pindell and Barrett, 1990).
esis in the northern Central Andes had started, probably This arc probably first encountered the Andes within
related to the same causal mechanisms that flipped the Ecuador because this was the boundary between volumi-
offshore arc to the north. Pindell (1993) suggested that nous and widespread plutonism and volcanism to the
this orogenesis was due to the early Aptian onset of the south (rapid subduction of normal Farallon plate) and
opening of the equatorial Atlantic (Pindell and Dewey, limited and localized volcanism or plutonism to the
1982; Pindell, 1985), which markedly accelerated South north (slow, shallow-dipping? subduction of the 12–25-
America westward across the mantle, thereby trans- km-thick Caribbean plate).
forming the Central Andes into a convergent arc in the Along the Andes, sands prograded into the foreland
sense of Dewey (1980). Peruvian orogenesis lightly area from both the east and west. Throughout the central
loaded the interior craton, creating additional accommo- Andes, marine accumulations became restricted to only
dation space and water depths for marine units such as locally transgressive areas. Farther south, where Atlantic
the Napo and Chonta formations. Evidence for compres- influences could be felt, marine conditions continued to
sion is found along the entire length of the Andes and prevail. A general decrease in Late Cretaceous subsi-
Amaime-Chaucha terrane, except perhaps in northern dence rates within the sub-Andean foredeep basins
Argentina, although the intensity varies considerably. could have been due to the waning of Peruvian compres-
The back-arc of Chile, including the Rocas Verdes basin, sional deformation.
began to close, and continental rift basins of central In the early Paleogene (Figure 7), westward drift of
Argentina were influenced by compression and foredeep South America caused the relative eastward advance of
deposition (Dalziel and Forsythe, 1985; Hallam et al., the Caribbean plate, which had a slight northward
1986). component of relative motion as well. The triple junction
The middle Cretaceous (Figure 5) marked the defined by the eastward continuation of the Panamanian
maximum extent of the seas across the craton, especially arc and the Andean trench thus migrated slowly
in the north where Andean trends intersected the northward, leading to much faster subduction of
passive margin trends. There, thermal subsidence of Farallon crust along the Andes in its wake. In turn, the
numerous Jurassic intracontinental and back-arc rifts volume of arc magmatism appears to have increased
combined to produce a broad, slowly subsiding epicon- northward, first in Ecuador and then into Colombia. By
tinental platform. The strand line was so far inland that early–middle Eocene time, crustal elements and accreted
deposition did not keep pace with subsidence. Pelagic material of the southeastern Caribbean plate (Caribbean
material accumulated at relatively high rates due to nappes) began to load the western and Maracaibo
upwelling-enhanced productivity to produce world portion of the northern margin’s shelf, creating a
class source rocks (e.g., La Luna Formation). Preserva- northward-deepening foredeep geometry (Trujillo and
tion was maintained by the intersection with the shelf Misoa formations) that triggered oil maturation along the
surface of the oxygen minima zone. In Ecuador and northern shelf of northwestern South America (Pindell,
Peru, foredeep subsidence allowed accumulation and 1991).

(text continues on p. 124)


106 Pindell and Tabbutt

Figure 1—Triassic–Early Jurassic (about 190 Ma) paleogeographic map of the Andes. Scale is
denoted by latitude and longitude ticks approximately 110 km apart.
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 107

Figure 1 (continued)
108 Pindell and Tabbutt

Figure 2—Middle–Late Jurassic (about 145 Ma) paleogeographic map of the Andes.
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 109

Figure 2 (continued)
110 Pindell and Tabbutt

Figure 3—Neocomian (about 125 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 111

Figure 3 (continued)
112 Pindell and Tabbutt

Figure 4—Aptian–Albian (about 105 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 113

Figure 4 (continued)
114 Pindell and Tabbutt

Figure 5—Cenomanian–Santonian (about 90 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 115

Figure 5 (continued)
116 Pindell and Tabbutt

Figure 6—Campanian–Maastrichtian (about 70 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 117

Figure 6 (continued)
118 Pindell and Tabbutt

Figure 7—Paleocene–Eocene (about 45 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 119

Figure 7 (continued)
120 Pindell and Tabbutt

Figure 8—Oligocene–middle Miocene (about 25 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 121

Figure 8 (continued)
122 Pindell and Tabbutt

Figure 9—Late Miocene–Recent (about 10 Ma) paleogeographic map of the Andes.


Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 123

Figure 9 (continued)
124 Pindell and Tabbutt

In the late Eocene–early Oligocene, between ano- assisted with east-west convergence in this area, leading
malies 13 and 18 (~33–38.5 Ma) (Cande and Kent, 1992), to more than 100 km of lateral tectonic escape by the
there was a marked (about twofold) westward accelera- roughly triangular Maracaibo block toward the free face
tion of South America across the mantle (e.g., Müller et of the Caribbean Sea. To the south, eastward migration
al., 1993). We suggest that this triggered the Incaic phase of the thrust front generally cannabalized the earlier
of Andean tectonism (intensification of the convergent foredeep and pushed the evolving foredeep axis craton-
arc) (Dewey, 1980), marked by eastward and westward ward. Numerous inter-Andean basins began to develop,
thrusting at the Andean flanks along most of the length especially in Peru and Bolivia, where the width of defor-
of the Andes. The chain was uplifted drastically at this mation was the greatest, possibly assisted by late
time, with erosion and deposition of considerable Paleozoic evaporitic décollement horizons. Segmentation
volumes of offlapping erosional materials in both the of the subducting Pacific plate further influenced
back-arc and fore-arc regions. In the southern Andes, magmatism and deformation style along strike, with flat-
compression and uplift were not as severe and the slab, nonmagmatic segments driving foreland deforma-
eastward extent of eroded molasse was limited. The tion in the Pampean and other ranges (Figure 9).
persistent epicontinental marine basins of southern
Argentina maintained contact with the Atlantic Ocean
throughout the Paleogene. HYDROCARBON CONSIDERATIONS
In the middle Tertiary (Figure 8), relative advance of
the Caribbean plate continued and more buoyant parts Settings for Source Rock Accumulation
of the Costa Rica–Panama arc began to collide, initiating and Preservation
Panama’s oroclinal geometry and eastward thrusting of
Colombia’s Central Cordillera over the Gualanday The paleogeographic maps (Figures 1–9) show that
foredeep basin. The Caribbean foredeep south of the five main Mesozoic–Cenozoic settings exist for source
Caribbean nappes was located in central Venezuela at rock deposition and preservation in the Andean system.
this point (La Pascua and Roblecito formations), with an These are indicated by formation names followed by a
east-west strike-slip regime developing between the two circled “s” symbol. The first setting is restricted rift
plates in the Falcon area north of Lake Maracaibo. A late basins with varying access to the sea during times of
Oligocene–early Miocene slowing of the rate of motion of back-arc extension. Examples include the Pucara and
South America relative to the mantle may have been Santiago formations, deposited in the Triassic–Jurassic
responsible for the period of generally finer grained back-arc basins of Peru and Ecuador (Figure 1). The
deposition (less convergent arc) in the Andean foreland second setting is thermally subsiding passive margin
basins, such as the middle Magdalena’s Mugrosa sections that developed during periods of slow sediment
Formation in the north (Figures 8, 9). Although the accumulation and high long-term relative sea level,
intensity of shortening for this period based on indi- where upwelling and the oxygen-minimum zone inter-
vidual accumulation rates seems less than for the Incaic sected the shelf. Examples are the middle Cretaceous La
phase, the Andes to the south continued to develop at Luna, Querecual, Gautier, and Villeta formations in the
this time into a continuous geographic barrier between north and the Lower Cretaceous Inoceramus Shale in the
marine systems of the fore-arc and the foreland. They south (Figures 4, 5). The third setting, related to the
provided a continuous source of molasse for the second but noteworthy on its own, is the rift structures
remainder of the Tertiary. that cross southern South America, aulacogens of the
Finally, in the middle Miocene–Recent (Figure 9), South Atlantic (e.g., San Jorge basin; Figures 2, 3), in
increased “Andean” orogenic uplift and development of which nonmarine and marine source rocks were
present-day relief (rejuvenation of convergent arc) deposited early on.
appears to have been driven by several factors: (1) The fourth setting is the tectonically downflexed
another, but less drastic, westward acceleration of South foredeep basins that formed east of the developing Andes
America across the mantle; (2) a progressive decrease in at times of high long-term eustatic sea level (during
the age of crust entering the trench along the length of Peruvian orogeny), when the epicontinental strand line
the Andes, causing resistance to subduction; and (3) was far to the east. Examples are the middle Cretaceous
rapidly increased rates of subduction of the Nazca plate Napo and Chonta formations of the north-central and
in the Miocene, which increased volcanism along the central Andes (Figure 5). The thick Devonian–Carbonif-
chain and may have thermally softened the belt, making erous source-prone section of Bolivia and southern Peru
it more deformable (Pitman et al., 1992). Erosion of the was probably also deposited in this tectonic setting
chain has produced massive volumes of Miocene–Recent during the sinistrally transpressive accretion of the
molassic detritus, creating thick foredeep sections that Arequipa massif (approximate location on Figure 1). The
extend well to the east and across most of the western fifth setting is along the Andean fore-arc at various times
fore-arc basins. in areas where terrigenous sedimentation was slow due
Eastward advance of the thrust belt has cannabalized to low Andean relief, vegetational retention, or dry
parts of the former (e.g., Incaic) foredeeps, thereby climate (Ziegler et al., 1981) and where upwelling and
creating the sub-Andes fold and thrust belt. In the north, other oceanographic factors presumably helped with
thrusting has jumped cratonward to the Eastern bioproductivity and maintainance of suboxic bottom
Cordillera and the Merida Andes of Colombia and conditions. A possible example is the Upper Cretaceous
Venezuela. The continued Panamanian collision has Redondo Shale of the Talara basin of Peru.
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 125

As for the marine source rocks in all these settings, The second source of sand was the uplifted and
both eustasy and tectonism helped to control source rock exposed thrust sheets in the deformation belts of the
deposition and preservation. Both combined to control Cordillera. The likelihood of these sandstones func-
accommodation space and water depth, while high tioning as reservoirs depends on two factors: (1) the
eustatic sea levels maintained a distant strand line in lithology of the strata being eroded and (2) the extent of
some settings during middle Cretaceous time and sorting that occurred during the transport of the detritus.
limited the input of sand. If the thrust sheets that were eroded did not contain
lithologies (e.g., quartz grains) conducive to maintaining
Classes of Sandstone Reservoir Deposition porosity upon redeposition and lithification, reservoir
potential will be low. Local variations in the exposure
Formations with known or potential reservoir can be very important in reservoir development. Since
potential are indicated in Figures 1–9 by a circled “r” much of the strata exposed in the Cordillera is either fine-
symbol following the formation name. Although produc- grained (carbonates, shales, or siltstones) or poorly
tion is achieved from some carbonate units, such as the sorted sandstones, the ability of the depositional system
latest Cretaceous Yacoraite Formation of the north- to enhance the sands by sorting is critical. In general,
western Argentine basin (Figure 6) and the Cogollo those sediments that were deposited proximal to the
Group limestones of the western Maracaibo basin thrust sheets tend to be poorly sorted and make poor
(Figure 4), quartzose sandstones provide most of the reservoirs, while those more distal to the region of defor-
Andean reservoirs. Sandstones came from four primary mation may be mature enough to serve as reservoirs.
provenances throughout the Phanerozoic: metamorphic Along the western side of the Andes, sandstone units
rocks of the Precambrian shield, reworked sedimentary within deltas that formed in the fore-arc region often
rocks of the Andean thrust belt, masses of crystalline function as reservoirs. These deltas were associated with
basement exposed in the thrust belt, and plutons and fluvial systems that drained the arc system or the interior
volcanoes of the magmatic arc. In the sub-Andean adjacent to the deformation front.
basins, the most important reservoir sandstones were The source of the third sandstone class is again from
derived from the metamorphosed, highly quartzose the Andes, but specifically from metamorphic or
Precambrian shield areas to the east of the Andean belt. plutonic basement core areas that do have a significant
Formation of fault-controlled rift basins which then quartz component. The Santander and Garzon massifs of
thermally subsided (Permian–early Mesozoic) was the mainly sedimentary eastern Cordillera of Colombia
followed by flexural deflection of the crust due to short- are examples (Figure 9). Although these areas are
ening and loading of early Andean thrust belts and sedi- presently exposed, they have not always been so. Thus,
mentary rocks (late Mesozoic–early Cenozoic). These their contribution of reservoir quality clastics is periodic,
processes created a “downhill” condition from the commonly contaminated by Andean volcanic and sedi-
shields to the back-arc and foredeep region (the present mentary sources. Additionally, massifs that are presently
eastern Andes). This broad, gentle slope allowed covered by sediments or volcanics may have been
westward-migrating fluvial sands to become clean, exposed and eroded during an earlier period. Uplift and
mature, and highly quartzose. Their point of final depo- exposure of the metamorphic terranes tend to coincide
sition and their westward extent was a function of the and persist after compressional deformation.
position of the strand line and paleoslope of the marginal The final sandstone class is the volcanogenic sandstone
seas. These were in turn dependent on eustatic sea level from the Andean magmatic rocks. Although most of the
and the intensity of orogenic development (downward plutonic rocks of the Andes are described in the literature
foredeep flexure or thermal uplift and subsidence) at the as “granitoids” or “granitic,” these are often intermediate
time. Thus, fluvial to coastal to shallow marine blanket- in silica content and only occasionally have high concen-
like sandstone deposits are common along the western trations of free quartz that can produce good quality
margins of the Precambrian shields and eastern flanks of reservoir sandstones. Thus, “sand” eroded from magmatic
flexural foredeep and thermally subsiding rift basin areas of the Andes does not usually produce quality
sections. reservoir sandstones. There are hypothetical exceptions,
Easterly derived sandstone horizons generally such as from areas where plutons may be tonalitic (inter-
coalesce with their vertical neighbors toward the east mediate but with free quartz crystals). But tonalites are
(i.e., shales pinch out to the east where basinal conditions rarely distinguished in the literature from “granitoids,”
were less developed). Likewise, shield-derived sand- which are typically dacitic or andesitic dominated by
stones pinch out toward the west away from the fluvial feldspars and pyroxenes that degrade into clays.
and coastal input of the eastern source area. This creates In the sub-Andean basins, there have been western
an interfingering of sandstones (reservoirs) and shale and eastern sources of sands, with the latter usually
(seals) that become sand-dominated to the east and producing the better quality reservoir sandstone (higher
shale-dominated to the west. In contrast, due to their quartz content). The region of interfingering of western
general narrow and elongate character and the influence and eastern sources of sandstone has been an important
of more local faulting, the facies of the early Mesozoic rift zone throughout the history of the basins for hydro-
basins tend to be less extensive and lithologic changes carbon potential. Depending on the type of basin
(pinching out of sandstones) are more abrupt than those flanking the Andes at various times, however, a wide
of the foredeep basins. swath of basinal deposits (marine and lacustrine shales,
126 Pindell and Tabbutt

limestones, evaporites, cherts, and fine clastics) often (addition of overburden) have been extremely intense
intervened. These are the sediment sequences that tend and rapid. In addition, subsequent events have affected
to have source rock potential. Such was the case during slightly different areas of the original autochthonous
at least two styles of tectonic history: back-arc rifting and depocenters, with relatively fresh source rock horizons
subsequent thermal subsidence (Triassic–Early Creta- quickly entering the oil maturation window during these
ceous) and foredeep basin flexure due to eastward over- successive periods of rapid tectonic and sedimentary
thrusting from the west (middle Cretaceous–early loading.
Cenozoic). In most regions along the present eastern Early Mesozoic rifts were sites of restricted environ-
Andes, eastward-migrating thrusts carrying sediments ment source rock deposition. High heat flow associated
originally deposited in the central basin have overridden with rifting may have driven early maturation in parts of
or interthrusted the quartz-rich easterly derived flanking these basins, perhaps into Early Cretaceous time, as
sandstones, so that updip migration from beneath or indicated in Colombia’s eastern Cordilleran back-arc
within the thrusts has tended toward the quartz sand- basin by detrital asphalt pebbles in the Hauterivian
stones. Prior to the Cenozoic, westerly derived sand- section (Campbell and Burgl, 1965) (Figure 3). Starting
stones from the Andes interfingered with the basinal first in the Peruvian and southern Ecuadorian region,
deposits and are in many places now metamorphosed middle Cretaceous onset of convergent arc behavior (late
and structurally interwoven with the intervening basinal Albian, Peruvian orogeny) helped to create the first
sedimentary sections. “foredeep” conditions, which were sites of additional
Locally during the Paleocene, and more commonly by source rock deposition (Napo and Chonta formations),
the Eocene, eastward migration of clastic sediment from due to tectonically and eustatically created accommoda-
the developing Andes mountains drowned out the tion space. These are correlative to the higher quality
eastern source such that clastic migration was eastward passive margin source rock units of Colombia, Vene-
away from the Andes, or at least to foreland fluvial trunk zuela, and Trinidad to the north.
systems that ran parallel to the Andes. At least part of the Maturation of all these Cretaceous source rock units,
cause of this drastic change was that late Eocene as well as the remaining source potential of the
compression and orogenesis (Incaic phase), and hence Triassic–Jurassic units, would become a function of
erosion, were sufficiently intense that the shear volume burial, hydrothermal effects, or structural thickening
of Andean clastics overwhelmed the component from during the subsequent successive tectonic pulses.
the eastern source. The line between clastic realms However, anomalous basement heat should be a factor
simply shifted eastward to the point where there was in areas that have undergone Cenozoic rifting,
essentially no topographic depression (foredeep) for volcanism, or ridge subduction. Examples of this type of
easterly derived sands to fill. This occurred despite the area include the central Andean Altiplano (Figures 8, 9),
fact that subsidence rates accelerated in most areas of the the arc axis, various strike-slip basins of the fore-arc, the
eastern foredeeps during this phase of Cenozoic deposi- Nazca–Antarctica–South America triple junction in Chile
tion. This condition led to a geometry in which the basal (Figure 9), and western Colombia, where an extinct ridge
surface of Cenozoic sandstones dips westward, while the system entered the trench from Oligocene to late
Eocene–Recent geomorphic surface has dipped generally Miocene time (Hardy, 1991) (Figure 9). In the sub-Andes
eastward. Thus, the latter Cenozoic “brown beds” form a basins, each successive advance of the thrust front and its
thick cover above the Cretaceous section, thickening associated molassic foredeep depocenter shifted the belt
toward their westerly source area. These Cenozoic of hydrocarbon maturation correspondingly eastward.
molassic beds are only locally important from a reservoir In the north, foreland thrusting with development of
standpoint because their porosity is limited and they lack foredeep depocenters occurred in the Campanian–Maas-
seals within them. Nevertheless, they are critical for trichtian (Umir-Colon and Mito Juan formations), in the
hydrocarbon systems in that they have played a major late Eocene–Oligocene Incaic phase (eastward thrusting
role in driving Cenozoic maturation by burial. Accurate and deposition of the Gualanday Group and equiva-
dating of these beds is critical for accurate prediction of lents), and in the middle(?) Miocene–Recent Andean
hydrocarbon maturation in many of the basins. phase (Neogene molasse units, uplift of eastern
Cordillera). The first phase probably did not initiate oil
Regional Maturation Mechanisms generation, but the second and third phases certainly
did. Incaic oil migration reached the Llanos basin, as the
Our assessment of Andean tectonics and sedimenta- eastern Cordillera was not yet elevated. In the central
tion allows general statements to be made about the Andes, the Peruvian phase may have triggered local
generation and accumulation of hydrocarbons. As previ- maturation, but the Incaic and Neogene phases were
ously outlined, the Andes have evolved in a series of stronger and fully capable of generating hydrocarbons.
tectonic pulses, mainly controlled by changes in plate In the southern Andes, evidence for Cretaceous compres-
motions, each of which has played some role in the sional deformation is meager, and Incaic deformation is
overall hydrocarbon history. That source rock units as relatively less than to the north, but Neogene phases
old as Ordovician and as young as Neogene have both were strong. At the extreme southern end of the chain,
become mature in the Neogene phase(s) of basin devel- the Rocas Verdes marginal basin was closed during the
opment attest to the significance of these pulses in middle Cretaceous and experienced moderate Incaic and
relation to hydrocarbons. This is because the Cenozoic Neogene phases of convergent deformation and
phases of orogenesis and associated basin subsidence orogenic sedimentation.
Mesozoic–Cenozoic Andean Paleogeography and Regional Controls on Hydrocarbon Systems 127

In addition to these primary phases of tectonism and reduces the volume of clastic material derived from
associated pulses of maturation and migration, the uplifted areas, particularly along parts of the western
original regional extent of source rock occurrence is flank of the Andes (Ziegler et al., 1981), with a corre-
another critical factor in assessing remaining potential. sponding reduction in burial maturation in the adjacent
For much of the eastern and sub-Andes belts, the swath basin. Despite such deviations, knowledge of the precise
of source rocks was fairly narrow, located in back-arc age of uplift and associated basinal subsidence can be
areas due to rifting or Cretaceous tectonic loading. In used to predict the onset of peak pulses of maturation.
contrast, toward the north in Colombia and western Migration routes and trapping can then be assessed by
Venezuela, the Cretaceous source rock depocenter was considering the paleogeography and structural configu-
far wider in the east-west dimension because of the more ration of the basin at the time. It should be noted,
regional occurrence of Jurassic lithospheric rift basins that however, that successive orogenic episodes are likely to
underwent thermal subsidence and marine transgression affect trapping potential and thermal conditions of struc-
in the Cretaceous. With each successive phase of tures developed during previous orogenic episodes.
tectonism and foredeep deposition, more and more of the There is a tendency for old oil generated during earlier
swath of source rocks was exhausted from west to east. orogenic phases to be remigrated, lost from the system,
In the Cenozoic, the earlier Incaic phase probably or driven to gas during subsequent phases of tectonic
triggered maturation of huge quanitities of oil because development.
few of the source rocks had been depleted by that time.
Subsequent phases not only had less source rock area to
mature but the quality of the remaining source rocks was CONCLUSIONS
poorer due to greater depositional proximity to the
shield. Therefore, along much of the Andes, Neogene Our analysis shows that tectonic, depositional, and
phases of tectonism have probably generated relatively hydrocarbon histories are closely interrelated, having
lesser volumes of oil. However, in the north, the original occurred in fairly discrete pulses through time, each with
source rock limit was sufficiently far east that the advent its own significance to hydrocarbon potential. Source
of Colombia’s eastern Cordillera thrust belt and foredeep rock units can be deposited during periods of rifting,
basin occurred well within the limits of excellent source passive margin sedimentation, and foredeep develop-
rocks. Hence, more Neogene oil was probably generated, ment during long-term sea level highstands, when the
although not necessarily trapped, in Colombia’s eastern strand line is far from the depocenter, and along the fore-
Cordilleran foothills and the Llanos basin than in more arc when clastic dilution is low and oceanic conditions
southward segments of the sub-Andes belts. In addition, are favorable. Quality reservoir units can be deposited,
the intermontane Maracaibo basin of western Venezuela depending on reworking and transport history, during
also underwent a major Neogene phase of generation cratonal transgression and regression and when quartz-
due to the uplift of and sedimentation from the Merida bearing blocks or strata in the Andes are eroded during
Andes, Santander massif, and Sierra Perijá. With respect uplift. Oil generation can be triggered by early rift-
to oil generation, the Neogene of the northern area is related heat flow, but more commonly by accumulation
more like the Eocene of more southward parts of the of sufficient overburden, either thrust wedges or
Andes. orogenic erosional products. In the sub-Andes, succes-
Another factor to consider that applies mainly to the sive eastward-advancing tectonic phases drove over-
southern Andes (northwestern Argentine basin and burden development which, where sufficient, triggered
southward; Figures 2, 3, 4) is that Jurassic–Cretaceous associated phases of hydrocarbon generation. The newly
rifts, associated with both the opening of the Atlantic and created potential of each successive phase in a given area
back-arc extension, cross the continent westward to at was dependent on the quality and distribution of source
least the sub-Andes belts (Figures 1–3). Early source rock rocks there. Neogene phase(s) generated large volumes
deposits in these basins represent “paths of oil potential” of oil from Mesozoic source rocks in the northern and
extending eastward from under the Andes foothills and southern Andes where the platformal and rift-bounded
foredeeps. The western ends of these paths have entered source rock units, respectively, extended well east of the
maturation as they were progressively overthrust and Incaic (late Eocene–Oligocene) foredeep basins. The
buried by the Andes and foredeep sediments. These middle Paleozoic source rock units of the central sub-
basins possess an important pulse of Neogene matura- Andes were also affected by eastward-stepping phases.
tion that may be in addition to an Incaic pulse. In areas In all areas, preservation of older oil was reduced by
between these rift basins, source rocks are of poorer Neogene destruction of preexisting (Incaic) structures,
quality and Neogene oil generation is therefore limited. which in many cases were once filled.
The considerations just discussed suggest an episod-
icity of oil generation in the Andes that is directly tied to
the episodicity of thrust belt development and foredeep Acknowledgments We thank John F. Dewey, Walter C.
deposition. This applies to both the sub-Andes basins Pitman, and Steve Cande for collaboration and assistance on
and the western fore-arc margin because episodes of various parts of our Andean study, and Sam Algar for helping
uplift will also drastically accelerate fore-arc sedimenta- to reduce our paleogeographic maps to the format presented
tion and burial maturation. This process is modified by here. Grants from Texaco and Mobil provided the initial means
climate to varying extents. For example, aridity greatly for much of this synthesis and are greatly appreciated.
128 Pindell and Tabbutt

REFERENCES CITED Müller, R. D., J.-Y. Royer, and L. A. Lawver, 1993, Revised
plate motions relative to the hotspots in the Atlantic and
Aspden, J. A., and W. J. McCourt, 1986. Mesozoic oceanic Indian oceans: Geology, v. 21, p. 275–278.
terrane in the central Andes of Colombia: Geology, v. 14, Pindell, J. L. 1985, Alleghanian reconstruction and the subse-
p. 415–418. quent evolution of the Gulf of Mexico, Bahamas and Proto-
Avé-Lallement, H. G., 1990. The Caribbean–South American Caribbean Sea: Tectonics, v. 4, p. 1–39.
plate boundary, Araya peninsula, eastern Venezuela: Pindell, J. L., 1991, Geologic rationale for hydrocarbon explo-
Twelve Caribbean Geologic Conference Transactions, St. ration in the Caribbean and adjacent regions: Journal of
Croix, U.S. Virgin Islands, p.461–471. Petroleum Geology, v. 14, p. 237–257.
Bourgois, J., J. Toussaint, H. Gonzalez, J. Azema, B. Calle et al., Pindell, J. L., 1993, Regional synopsis of Gulf of Mexico and
1987. Geologic history of the Cretaceous ophiolitic Caribbean evolution, in J. L. Pindell and R. Perkins, eds.,
complexes of northwestern South America (Colombian Mesozoic and early Cenozoic development of the Gulf of
Andes): Tectonophysics, v. 143, p. 307–327. Mexico and Caribbean region: GCSSEPM Foundation, 13th
Campbell, C. J., and H. Burgl, 1965, Section through the Annual Research Conference Proceedings, Houston, Texas,
eastern Cordillera of Colombia, South America: GSA p. 251–274.
Bulletin, v. 76, p. 567–590. Pindell, J. L., and S. F. Barrett, 1990, Geologic evolution of the
Cande, S. C., and D. V. Kent, 1992, A new geomagnetic Caribbean region: a plate tectonic perspective, in G. Dengo
polarity time scale for the Late Cretaceous and Cenozoic: and J. E. Case, eds., Decade of North American Geology, v.
Journal of Geophysical Research, v. 97, p. 13917–13951. H, The Caribbean Region: GSA, p. 405–432.
Case, J. E., T. L. Holcombe, and R. G. Martin, 1984, Map of Pindell, J. L., and J. F. Dewey, 1982, Permo-Triassic reconstruc-
geologic provinces in the Caribbean region, in W. E. Bonini, tion of western Pangea and the evolution of the Gulf of
R. B. Hargraves, and R. Shagam, eds., The Caribbean–South Mexico/Caribbean region: Tectonics, v. 1, p. 179–212.
American plate boundary and regional tectonics: GSA Pindell, J. L., and J. P. Erikson, 1994, The Mesozoic passive
Memoir 162, p. 1–31. margin of northern South America, in J. A. Salfity, ed.,
Daly, M. C., 1989, Correlations between Nazca/Farallon plate Cretaceous Tectonics of the Andes: Braunschweig/
kinematics and fore-arc basin evolution in Ecuador: Weisbaden, Earth Evolution Sciences, Vieweg Publishing,
Tectonics, v. 8, p. 769–790. p. 1–60.
Dalziel, I. W. D., 1986, Collision and cordilleran orogenesis: an Pindell, J. L., K. D. Tabbutt, J. F. Dewey, W. C. Pitman III, and
Andean perspective, in M. P. Coward and A. C. Ries, eds., S. Cande, 1992, Phanerozoic geologic evolution of the
Collision tectonics: GSA Special Publication 19, p. 389–404. Andean system and implications for hydrocarbon and
Dalziel, I. W. D., and R. D. Forsythe, 1985, Andean evolution resource potential: Lyme, New Hampshire, Tectonic
and the terrane concept, in D. G. Howell, ed., Tectonostrati- Analysis, 1028 p., 32 plates.
graphic terranes of the circum-Pacific region: Circum- Scotese, C. R., J. Y. Royer, R. D. Müller, et al., 1987, Atlas of
Pacific Council for Energy and Mineral Resources, Mesozoic and Cenozoic plate tectonic reconstructions: Pale-
p. 565–581. ogeographic Mapping Project, Institute for Geophysics,
Dewey, J. F., 1980, Episodicity, sequence and style at conver- University of Texas, Austin, Technical Report 90, 31 p.
gent plate boundaries, in D. W. Strangeway, ed., The conti- Stephan, J. F., C. Beck, A. Bellizzia, and R. Blanchet, 1980, La
nental crust and its mineral deposits: Geologic Association chaîne caraîbe du Pacifique à l’Atlantique: Memoires
of Canada Special Paper 20, p. 553–574. Bureau Recherches Geologiques et Minieres, v. 115,
Dewey, J. F., and J. L. Pindell, 1985, Neogene block tectonics of p. 38–59.
eastern Turkey and northern South America: continental Tschanz, C. M., R. F. Marvin, J. Cruz, H. H. Mehnert, and G. T.
applications of the finite difference method: Tectonics, v. 4, Cebula, 1974, Geologic evolution of the Sierra Nevada de
p. 71–83. Santa Marta, northeastern Colombia: GSA Bulletin, v. 85, p.
Dewey, J. F., and J. L. Pindell, 1986, Neogene block tectonics of 273–284.
eastern Turkey and northern South America: continental Ziegler, A. M., S. F. Barrett, and C. R. Scotese, 1981, Paleocli-
applications of the finite difference method: Reply: mate, sedimentation, and continental accretion: Philisoph-
Tectonics, v. 5, p. 703–705. ical Transactions of the Royal Society of London, v. A301,
De Wit, M. J., 1977, The evolution of the Scotian arc as the key p. 253–264.
to the reconstruction of southwestern Gondwanaland: Ziegler, A. M., D. B. Rowley, A. L. Lottes, D. L. Sahagian, M. L.
Tectonophysics, v. 37, p. 53–81. Hulver, and T. C. Gierlowski, 1985, Paleographic interpre-
Grosser, J. R., 1989, Geotectonic evolution of the western tation: with an example from the mid-Cretaceous: Annual
Cordillera of Colombia: new aspects from geochemical data Reviews of Earth and Planetary Science, v. 13, p. 385–425.
on volcanic rocks: Journal of South American Earth
Sciences, v. 2, p. 359–369.
Hallam, A., L. Biro-Bagoczky, and E. Perez, 1986, Facies Authors’ Mailing Addresses
analysis of the Los Valdes Formation (Tithonian–Hauteri- James L. Pindell
vian) of the high Cordillera of central Chile, and the paleo- Tectonic Analysis Inc.
geographic evolution of the Andean basin: Geologic
P.O. Box 87, One Lyme Common
Magazine, v. 132, p. 425–435.
Hardy, N. C., 1991, Tectonic evolution of the easternmost Lyme, New Hampshire 03768
Panama basin: some new data and inferences: Journal of U.S.A.
South American Earth Sciences, v. 4, p. 261–269.
MacDonald, W. D., B. L. Doolan, and U. G. Cordani, 1971, Kenneth D. Tabbutt
Cretaceous–early Tertiary metamorphic K/Ar age values Evergreen State College
from the southern Caribbean: GSA Bulletin, v. 82, Olympia, Washington 98505
p. 1381–1388. U.S.A.
Paleozoic Basins

VOLCAN LANIN in the Cordillera Patagónica, Neuquén province,


Argentina. The Holocene basaltic cone of Lanín (3776 m) owes its elevation to
the support of the underlying Faja Batolítica, a Permian–Triassic granodioritic
basement.

Edgar Ortiz, 1994, watercolor, 30 × 23 cm


Hydrocarbon-Bearing Late Paleozoic Glaciated Basins
of Southern and Central South America

N. Eyles A. B. França
Department of Geology Nexpar, Petroleo Brasileiro S.A.
University of Toronto Curitiba, Brazil
Toronto, Ontario, Canada
C. H. Eyles
G. Gonzalez Bonorino Department of Geography
Department of Geology McMaster University
Conicet–University of Buenos Aires Hamilton, Ontario, Canada
Buenos Aires, Argentina

O. López Paulsen
YPFB, Gerencia de Exploración
Santa Cruz, Bolivia

Abstract

A lthough glaciated basins are usually associated with nonproductive, poorly sorted strata, hydrocarbons
occur in several late Paleozoic glaciated basins of central and southern South America. In Bolivia, the
Chaco-Tarija basin has commercial production from more than 30 fields in glacially influenced submarine
channel systems (Palmar, Santa Cruz, and Bermejo fields) that accounts for about 60% of current national
reserves. Correlative deposits in Argentina host the Campo Durán and Madrejones oil fields. In Brazil, the
Paraná basin has significant but as yet subcommercial gas shows in thick marine turbidite sandstones of the
glacially influenced Itararé Group. The Chaco-Paraná basin of Argentina is one of the largest onshore targets
for exploration in South America, but it is virtually untested.
Glacially influenced foreland basins of Argentina (Tepuel and Paganzo-Maliman) contain complex
glacigenic stratigraphies of interbedded tillites and poorly prospective sandstones. In contrast, the glacially
influenced marine infills of intracratonic basins in Brazil (Paraná), Bolivia, and Argentina (Chaco-Tarija and
Chaco-Paraná) contain thick sequences of pebbly mudstones and regionally extensive reservoir quality sand-
stones. The key to the occurrence of good reservoirs and associated trapping mechanisms in these intracratonic
basins is the interplay of sediment supply, regional tectonics, and relative sea level changes. Glacial scouring of
extensive cratons by ice sheets resulted in the delivery of huge volumes of glaciofluvial sand to deltas. Struc-
tural control of drainage patterns on the craton by basement lineaments resulted in persistent sediment sources
and depocenters. Frequent earthquake activity along reactivated basement lineaments resulted in downslope
mass flow of deltaic sediments and the deposition of thick, amalgamated sand turbidites (reservoirs). Pebbly
mudstone seals most likely record higher relative sea levels, resulting from basin subsidence, and deposition
from suspended sediment plumes and icebergs. Source rocks are provided by Devonian and Permian shales.
This model may be applicable to other parts of Gondwana that contain thick, prospective sandstones in
glacially influenced intracratonic basins.

Resumen

L as cuencas glaciares estan usualmente asociadas a estratos pobremente escogidos y carentes de atractivo
comercial desde el punto de vista de hidrocarburos, sin embargo en varias cuencas glaciares del Paleo-
zoico tardio en el sur y centro de Suramerica, se ha probado la existencia de petroleo. En Bolivia, la cuenca
Charco-Tarija tiene produccion comercial en mas de 30 campos ubicados en sistemas de canales submarinos
con influencia glacial (e.j., campos Palmar, Santa Cruz, y Barmejo), que representan aproximadamente el 60%
de las reservas nacionales. Depositos correlativos en Argentina corresponden a los campos petroliferos de
Duran y Madrejones. En Brasil, la cuenca Paraná tiene cantidades significativas de gas (en condiciones
subcomerciales) en las espesas areniscas marinas turbiditicas con influencia glacial del grupo Itarare. La
Cuenca Chaco-Paraná de Argentina es uno de los objetivos costeros mas grandes para la exploracion en
Suramerica pero virtualemente no ha sido probada.

Eyles, N., G. Gonzalez Bonorino, A. B. França, C. H. Eyles, and O. López Paulsen, 1995, 165
Hydrocarbon-bearing late Paleozoic glaciated basins of southern and central South
America, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 165–183.
166 Eyles et al.

Las cuencas foreland con influencia glacial ubicadas en Argentina (Tepuel y Paganzo-Maliman), contienen
complejas estratigrafias glaciogenicas de tilitas interestratificadas con areniscas pobremente prospectivas. En
contraste, los rellenos marinos con influencia glacial de las cuencas intracratonicas de Brasil (Paraná), Bolivia y
Argentina (Chaco-Tarija), Charco-Paraná), contienen gruesas secuencias de lodolitas gravosas (diamictitas,
sellos) y areniscas con calidad de yacimientos de extension regional. La clave para la formacion de buenos
yaciementos y mecanismos de entrampamiento asociados en estas cuencas intracratonica es las interaccion
entre el aporte de sedimentos, la tectonica regional y los cambios relativos en el nivel del mar. La socavamienta
glacial de extensos cratones por la accione las acumulaciones de hielo, resulta en el transporte de volumenes
muy grandes de arenas glaciofluviales hacia los deltas: el control estructural sobre los patrones de drenaje del
craton, ejercido por las alineaciones en el basamento, resulta en fuentes de sedimento y depocentros persis-
tentes. La actividad frecuente de terremotos a lo largo de las alineaciones reactivadas del basamento, ocasionan
flujos masivos descendentes de sedimentos deltaicos y la sedimentacion de grueasa aremas turbiditicas amal-
gamadas (yaciementos). Los lodolitas gravosas que actuan como sellos muy problemente registran niveles del
mar relativamente mas altos, resultantes de la subsidencia de la cuenca, y de la depositacion por corrientes de
sedimentos suspendidos y tempanos de hielo. Las rocas generadoras son lutitas de edad Devonico y Permico.
Este modelo puede ser aplicable a otras partes de Gondwana que contiene areniscas espesas y prosectivas en
cuencas intracratonicas con influencia glacial.

INTRODUCTION depositional environments and the stratigraphy of


glaciated basins are first discussed, stressing the signifi-
Late Paleozoic glacioclastic strata occur in many sedi- cance of regional tectonics and the selective preservation
mentary basins in South America, southern Africa, India, of glacially infuenced marine sediments. Reworking of
the Arabian Peninsula, Australia, Antarctica, and glacioclastic sediment in marine settings is the key to the
Malaysia (Hambrey and Harland, 1981). Glaciation of the occurrence of hydrocarbons in the late Paleozoic
southern continents between about 360 and 250 Ma glaciated basins discussed here.
coincided with the greatest episode of coal deposition in The Chaco-Tarija basin is the most important since it
earth’s history (Langford, 1991). Late Paleozoic glacio- has commercial production in Bolivia and Argentina.
clastic strata are associated with significant hydrocarbon Glacially influenced strata of the Bermejo, Palmar, and
resources in Australia (Youngs, 1975; Williams et al., Santa Cruz fields are important oil producers in Bolivia,
1985; Goldstein, 1989; Redfern, 1991), Oman (Levell et al., where about 60% of current reserves occur in glacially
1988), and South America (Montes de Oca, 1989; França influenced deposits. Correlative deposits in Argentina
and Potter, 1991). Many basins, particularly in Australia host the Tranquitas, Campo Durán, Icua, and Madre-
and parts of South America, contain thick, prospective jones oil fields and about 80% of national reserves. The
sandstone intervals in otherwise poorly sorted glacially Paraná basin in Brazil has significant but as yet subcom-
influenced strata (e.g., O’Brien et al., 1992). The origin of mercial gas shows in the glacigenic Itararé Group, and
these sandstones is not well understood. important coal deposits also occur in early postglacial
In South America, the presence of late Paleozoic glacial strata of the Rio Bonito Formation. The Chaco-Paraná
deposits was first recognized in the closing decades of the basin of northeastern Argentina is one of the largest
nineteenth century. Extensive outcrops in Brazil were the onshore targets for exploration in South America that is
first to be documented (Derby, 1888; Coleman, 1926), and virtually untested. In western Argentina, the Paganzo-
late Paleozoic glacial horizons were of central importance Maliman and Tepuel basins are nonproductive, high risk
in Du Toit’s (1927) and Wegener’s (1929) correlations areas but are of considerable interest because their
across the South Atlantic. Details of the glacial record in extensive outcrops yield valuable information about
South America and elsewhere in Gondwana have been facies and geometries of glacial deposits in the subsurface.
presented by Frakes and Crowell (1969), Crowell (1983),
Caputo and Crowell (1985), Dickins (1985, 1993), and
contributors to the compilation of Hambrey and Harland
(1981). Renewed economic interest in hydrocarbons in
glaciated basins in Argentina, Bolivia, Paraguay, and
SELECTIVE PRESERVATION OF
Brazil has promoted the study of facies, sequence strati- GLACIALLY INFLUENCED MARINE
graphy, and age relationships across the continent. A SEDIMENTS IN GLACIATED BASINS
major focus of this work is identification of glacial deposi-
tional environments, better understanding of tectonic Ancient glacial strata are commonly interpreted in
setting, and closer resolution of the controls on deposi- terms of what is understood of deposits at the margins of
tional sequences, such as climate and eustatic sea level modern and Pleistocene continental ice masses. These
changes. Unfortunately, many data are proprietary and deposits are usually dominated by poorly sorted, often
there is a dearth of published literature. bouldery sedimentary facies referred to as “tills” and are
This paper reviews what is currently understood of associated with complexly structured ice contact strati-
the depositional fills in late Paleozoic glaciated basins in graphies and glaciotectonically deformed substrates
southern and central South America and relates this to recording the subglacial bulldozing and deformation of
exploration for hydrocarbons. Selected aspects of glacial preexisting strata (e.g., Brodzikowski and Van Loon,
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 167

1991; Eyles and Eyles, 1992). Correspondingly, pre-Pleis-


tocene stratigraphic intervals containing multiple tillite
horizons are often interpreted in terms of climatically
driven advances and retreats of grounded terrestrial ice
margins (Hambrey and Harland, 1981).
Other investigators of earth’s long glacial record
instead stress a history of selective preservation of
glacially influenced strata in marine basins (e.g., Frakes
and Crowell, 1969). Using Pleistocene basins as analogs,
Eyles (1993) estimated that less than 6% by volume of the
pre-Pleistocene glacigenic rock record consists of conti-
nental glacial deposits. The primary record of glaciation
is stored in marine basins that developed in a wide range
of tectonic settings adjacent to glaciated areas. The direct
role of ice was limited in most cases to the areal scour of
continental surfaces and the production and delivery of
considerable volumes of glaciofluvial sand and mud to
basin margins where it was then reworked and depo-
sited by normal marine processes. The term glacially
influenced has been used to describe such basins (Eyles et
al., 1985) in recognition of the indirect depositional role
played by glaciers. Basin fills are distinguished by
marine strata dominated by thick sandstones associated
with complexes of pebbly mudstone (diamictite) facies,
emplaced by ice rafting of coarse debris into glacially
derived marine muds (“rain-out” diamictite) and re-
deposition as debris flows. These ideas are particularly
appropriate to understanding the infills of late Paleozoic Figure 1—Late Paleozoic sedimentary basins discussed in
glaciated basins in central and southern South America. this paper. Superimposed are dominant detrital sources
and paleocurrents (arrows). Note structural compartmen-
talization of the intracratonic Chaco-Tarija, Paraná, and
Chaco-Paraná basins by continent-crossing late Protero-
LATE PALEOZOIC GLACIAL RECORD IN zoic lineaments, arches, terrane boundaries, and Trans-
SOUTH AMERICA brasiliano fold belt.

Caputo and Crowell (1985) showed that the timing of


late Paleozoic glaciation across Gondwana is diachro- orogeny. In addition, the presence in polar Gondwana of
nous and proposed that this could be explained by the emergent landmasses presently underlying Patagonia
migration of Gondwana across the south pole. Veevers and the Antarctic Peninsula, either through collisional
and Powell (1987) argued that continental-scale glacia- accretion (Ramos, 1989) or tectonic uplift during the
tion was initiated by strong Namurian uplift along the Chanica orogeny, may have had the effect of bringing
western collisional margin of South America. They extensive land areas into high latitudes and thus fostering
envisaged separate ice centers located on pericratonic regional glaciation. The first definite record of glaciation
mountain ranges. Recent data suggest, however, that in southwestern Gondwana appears about Tournaisian
glaciation in Andean South America started as early as time in the northern Chaco-Tarija basin of Bolivia and
the latest Devonian (Diaz and Lema, 1991; Grahn and Peru (Diaz and Lema, 1991; López Paulsen et al., 1992).
Caputo, 1992) and that the principal ice centers were Older glacioclastic strata of Late Ordovician–Early
located not along the active plate margin but inboard Silurian (early late Llandovery–earliest Wenlockian) time
along interior basement highs, such as the Sierras are present in Bolivia and Brazil and are attributed to the
Pampeanas and North Patagonia massif (Figure 1) polar positioning of Gondwana (see Grahn and Caputo,
(González Bonorino, 1991). A major complication is that 1992).
the different ages of glacial strata from basin to basin Subsequent late Paleozoic glaciation of southern South
may reflect not only possible migration of ice centers and America affected foreland basins in western Argentina
changing climate across the continent but also the (Tepuel and Paganzo-Maliman) and intracratonic basins in
relative timing of basin subsidence and sediment preser- eastern Bolivia, Brazil, and Paraguay (Paraná and Chaco-
vation. Paraná; Figure 1). Following the Chanica orogeny, subsi-
The onset of glaciation in central and southern South dence and deposition started earlier (Early Carbonif-
America in general appears to have coincided with conti- erous) in the western basins than in the eastern basins,
nental uplift along the Pacific margin of Gondwana where preservation of a glacial record did not begin until
during the Late Devonian–Early Carboniferous Chanica the Late Carboniferous (Figure 2).
168 Eyles et al.

Figure 2—Stratigraphic correla-


tion chart for basins discussed
in this paper based on recent
biostratigraphy and the identifi-
cation of major transgressions
(indicated by wedges). Conflict
exists over the age of the
Tupambi Formation, which
some workers consider entirely
Pennsylvanian (Late Carbonif-
erous). On the right, the
approximate time of two major
diastrophic and orogenic
phases are shown, separated
by protracted extensional
phases that favored preserva-
tion of glacial deposits.

Tectonic Controls on Glacially Influenced developed into a large ice sheet, perhaps similar in extent
Sedimentation and Hydrocarbon Potential to that on modern Antarctica (González Bonorino, 1992).
The sedimentary record of this ice sheet has been
All the basins considered here, with the possible preserved on shallow marine foreland basins such as the
exception of Tepuel, had a protracted early Paleozoic Paganzo-Maliman and Tepuel basins of Argentina.
sedimentation history that was interrupted by Chanica These areas contain complex and poorly sorted glacial
compression and uplift of the continent, the withdrawal deposits with limited potential for hydrocarbons.
of epeiric Devonian seas, and widespread subaerial and In contrast, within the interior of southwestern
shallow marine erosion. Along the paleo-Pacific margin Gondwana, widespread erosion below a continental-
of Gondwana, early Paleozoic continental margin scale ice cover resulted in the development of a strati-
deposits were thrust onto the foreland, which subsided, graphic gap spanning the Late Devonian–Late Carbonif-
allowing Early Carboniferous transgression and sedi- erous. These conditions were terminated when
mentation in the Paganzo-Maliman and Tepuel basins. tectonically driven subsidence, reflecting changing
The hinterland remained largely submerged until the stresses along the compressional and transpressional
late Early Permian San Rafael diastrophic phase (Fig- Laurussian margin of Gondwana (Figure 3), resulted in
ure 2). At this time, the major influx of sediment was basin subsidence, marine incursion, development of
derived from crystalline terranes in the Pampeanas and separate ice centers, and deposition of glacially influ-
Patagonia highlands (Figure 2) (González Bonorino, enced sediments in the intracratonic Chaco-Tarija,
1991). Subsequent intracratonic subsidence within the Paraná, and Chaco-Paraná basins (Figure 2). The record
continental interior, possibly in response to the clockwise of direct sedimentation by ice is restricted in these basins
rotation of Gondwana against Eurasia (see Figure 3), led and is preserved in marine strata.
to renewed flooding of the Paraná and Chaco-Paraná
basins. These intracratonic basins were strongly influ-
enced by basement faulting along late Proterozoic linea- FORELAND BASINS
ments (Tankard et al., 1995). They received detritus from
surrounding shield areas, including those of southern Paganzo-Maliman Basin of Argentina
Africa and marginal uplifts underlain by Proterozoic–
lower Paleozoic crystalline and sedimentary rocks The Paganzo-Maliman basin contains up to 3 km of
(Figure 1) (Frakes and Crowell, 1972). Lower Carboniferous–middle Permian strata in central
Starting possibly in the latest Devonian or earliest western Argentina, cropping out in the Cuyo pre-
Carboniferous, ice caps grew on the elevated interior of Cordillera foreland thrust belt and in the block-faulted
Gondwana and by the early Late Carboniferous had Sierras Pampeanas (Figures 1, 4a). Strata in the central
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 169

Figure 3—Early Mississip-


pian–Pennsylvanian
(Early–Late Carboniferous)
clockwise rotation of
Gondwana against Laurussia
(after Leighton and Kolata,
1990), and subsidence of
intracratonic basins allowing
preservation of a glacially
influenced marine strati-
graphic interval containing
prospective reservoir-quality
sandstones (see Figures 7, 8).
Extent of ice cover is
schematic and approximate.

Figure 4—Stratigraphic
framework for (a) Paganzo-
Maliman and (b) Tepuel basins
in depth-distance plots. For
clarity, Andean deformation
has been removed, but struc-
tural shortening, which could
extend the Paganzo-Maliman
section by 60–90 km, has not
been compensated for. Inset
shows an outcrop map for the
Tepuel basin (basin boundary
is dashed line) and major
roads. Precambrian basement
is exposed in the east and
north. Stratigraphic section
goes from Esquel to Cerro
Catreleo. Abbreviations: C,
Cerro Catreleo; E, Esquel; Tk,
Tecka Range; Tp, Tepuel
Range; NPM, northern
Patagonia massif.
170 Eyles et al.

and eastern parts of the basin constitute the Guandacol, (González, 1981; López Gamundi, 1983; Milana and
Tupe, and Patquia formations and those in the west Bercowski, 1987). In the Hoyada Verde Formation, rain-
occur in the Hoyada Verde, Maliman, and other forma- out diamictites are capped by a boulder pavement that
tions (Figure 2). Stratigraphic details are summarized in records winnowing of the diamictites and are blanketed
Archangelsky et al. (1987), Azcuy et al. (1987), López by basinal mudstones. Shallow marine and terrestrial
Gamundi et al. (1987), and González Bonorino (1991). beds in the Tres Saltos Formation complete the sequence
The basin fill essentially comprises one major unconfor- (Figures 3, 5a).
mity-bounded unit that onlaps a tilted and erosionally
beveled lower Paleozoic substrate and oversteps crys- Tepuel Basin
talline basement in the Sierrras Pampeanas (Figure 4a).
The unit is divisible into three subunits: lower (Lower The Tepuel basin contains as much as 5 km of Lower
Carboniferous, which includes the El Paso and Maliman Carboniferous–Lower Permian sedimentary rocks
formations), middle (Upper Carboniferous, approximately (Tepuel Group and Jaramillo, Pampa de Tepuel, Rio
Namurian–Westphalian and which includes the Hoyada Genoa formations) (Figures 2, 4b) that are well exposed
Verde, Guandacol, and Tupe formations), and upper in the sub-Andean foothills. Much of the Carboniferous
(Upper Carboniferous, approximately Stephanian–Lower dips into the subsurface south of the Tepuel Ranges so
Permian, which includes the Tres Saltos and Patquia that the southern half of the basin exposes only latest
formations). Overall, nonmarine deposits intertongue Carboniferous–Lower Permian sections (Figure 4b,
westward with shallow marine strata (Figure 4a). Western inset). Cenozoic compression, superimposed on
sections expose deltaic and offshore strata (El Paso and Mesozoic extension, resulted in block faulting and gentle
Hoyada Verde formations) that are overlain by littoral and folding of the Tepuel Group strata over much of the
terrestrial rocks (Tres Saltos Formation). Eastern sections basin (Andreis et al., 1987; Archangelsky et al., 1987;
show an upward passage from lacustrine (Guandacol González Bonorino, 1991).
Formation) to alluvial plain (Tupe Formation) to eolian The Tepuel Group was deposited on a westward-
(Patquia Formation) strata (Figures 2, 4a). Glacial and facing shallow marine shelf and is dominated by wave-
glacially influenced deposits are mostly restricted to the worked sandstones interbedded with bioturbated
middle subunit in two major settings: lacustrine in the east mudstones (Figure 4b) (Page et al., 1984; González
and coastal plain and estuarine to open marine shelf in the Bonorino and González Bonorino, 1988). Petrographic
west (these facies are described later). Dropstones occur data reveal a persistent source area in the Patagonia
locally in the upper subunit, but may reflect seasonal ice highlands, then underlain by crystalline rocks similar to
cover in lakes. González (1990) has reported lower those of the North Patagonia massif (Figure 4b, inset)
Carboniferous tillites and dropstones in the area, but this (Frakes and Crowell, 1969; González Bonorino, 1992).
requires confirmation. Higher subsidence rates in the Tepuel compared to the
Paganzo-Miliman basin favored the preservation of a
thicker marine succession. Because of the more open
Sedimentary Facies
marine setting, there was greater opportunity for
In the Early Carboniferous, a wide coastal plain reworking of glacioclastic sediment into shallow marine
connected the Pampeanas highlands with a western shelf sand bodies across the postglacial shelf.
margin. In the early Late Carboniferous, an ice cap
centered on the Pampeanas highlands episodically Sedimentary Facies
expanded onto the coastal plain and shelf. Upper The lower part of the Pampa de Tepuel Formation
Carboniferous lacustrine deposits in the eastern Cuyo contains three intervals of glacial and glacially influenced
pre-Cordillera delineate a foreland lacustrine system that deposits (G1, G2, and G3) (Figure 5). Chaotically bedded
occupied the Zonda-Villicum trough (Figure 4a). The diamictites are characterized by exotic striated boulders
trough must have been at least sporadically connected to of granite in an argillaceous sandstone matrix and rest on
the open sea during highstands to account for fossil- severely deformed shelf deposits. These beds are most
iferous marine intercalations. Overall, the trough fill easily interpreted as mass flows accompanying the
constitutes a coarsening-upward deltaic succession downslope movement and deformation of underlying
(Guandacol Formation) with thin diamictites containing sediment, although a direct glacial origin as tillites cannot
striated clasts resting on supposedly glacially striated as yet be discounted. Overlying massive diamictites, up to
pavements (Bossi and Andreis, 1985). Diamictites are 60 m thick, show clasts dispersed in a mudstone matrix
intercalated with mudstone intervals containing drop- that is transitional to shelf mudstones and are interpreted
stones and slump structures and may represent the as rain-out deposits. These facies are truncated by well-
downslope mass flow of glacioclastic debris in a fan delta sorted wave-worked sandstone bodies up to 10 m thick
setting. These strata are poorly prospective because sand- and 2 km long, elongated in the direction of dominant
stones are limited to occasional, thin (1 m) storm deposits sediment transport (González Bonorino et al., 1988). Slatt
reworked from underlying glacial sediment. (1984) has described the development of Holocene sand
In the western part of the basin, sedimentary succes- bodies on the outer Canadian Atlantic shelf as a result of
sions are also diamictite dominated and poorly prospec- postglacial reworking of glacial sediments. The same
tive. Diamictites rest on deformed early Paleozoic model can be applied to the sandstone bodies of the
substrates and are overlain by thick rain-out diamictites Tepuel basin.
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 171

1989). Devonian and Permian shales of the Ponta Grossa


and Irati formations, respectively, are source rocks for
hydrocarbons in sandstones of the Itararé Group. Core
and downhole geophysical data are available from about
100 wells, but the collection of high-resolution seismic
studies is prevented by a thick cover of Late Jurassic
flood basalts (França and Potter, 1991). The Itararé Group
spans the latest Westphalian–earliest Stephanian (about
300 Ma) to latest Artinskian–earliest Kungurian (about
260 Ma) and is comparable in age to the glacigenic
Dwyka Formation of the Karoo basin in southern Africa
(Visser, 1990). Marine microflora are present throughout
the Itararé Group, including the genera Baltisphaeridium,
Micrhystridium, Veryhachium, Leiosphaeridia, Navifusa,
Cymatiosphaera, and Tasmanites (Daemon et al., 1992).
Three formations (Lagoe Azul, Campo Mourao, and
Taciba) can be recognized within the Itararé Group
(Figure 6), each overstepping the underlying formation
and recording renewed basin subsidence by faulting
along steeply dipping basement lineaments (see Eyles et
al., 1993). Subsidence and expansion of the basin was
strongly asymmetric to either side of a broad, northwest-
trending structural belt along the trend of the Guapiara
lineament, which is interpreted as a major intraplate
boundary in South America (Eyles and Eyles, 1993;
Tankard et al., 1995).

Sedimentary Facies
Each formation of the Itararé Group is composed of a
lowermost member of thickly bedded, amalgamated
turbidite sandstones (e.g., Cuiaba Paulista Member)
overlain by a fine-grained member dominated by shales,
muddy debris flows, and rain-out diamictites (e.g.,
Chapeu do Sol Member) (Figures 6, 7, 8).
The most striking characteristics of the sandstones in
the Itararé Group are their thickness, massive character,
and textural homogeneity. The sandstones are predomi-
nantly fine to medium in texture and are lithologically
immature, containing large quantities of lithic fragments.
A well-defined secondary porosity can be identified
(Figure 9). The sandstones consist of massive, graded,
Figure 5—Stratigraphic column for the middle part of the and deformed facies emplaced by turbidity currents.
Tepuel Group in the Tecka Range. G1–G3 are prominent Homogeneous sandstone bodies are as thick as 200 m,
glacial intervals; see text for details. See Figure 4 for
suggesting that they are the product of the repeated
location.
stacking and amalgamation of Bouma A turbidite beds.
These facies (e.g., Cuiaba Paulista Member) have a well-
INTRACRATONIC BASINS defined cylindrical gamma ray signature and give rise to
remarkably consistent dipmeter patterns (Figure 8a).
Paraná Basin of Brazil Chaotically deformed facies are locally present and
record postdepositional downslope slumping; liquefac-
The Paraná basin of southern Brazil is the largest tion structures are common. Other thick sandstone
(1.6 × 106 km2) late Paleozoic basin in southern South bodies (e.g., Campo Mourao Member) show well-defined
America (Figure 2). It contains a thick (1400-m) predomi- fining-upward sequences 10–30 m thick composed of
nantly marine glacial succession, the Itararé Group (dos graded sandstones facies (Figure 8b). These facies
Santos, 1987; França and Potter 1991), which has strong contain appreciable muddy matrix material that severely
similarities to strata in the Chaco-Tarija basin of Bolivia restricts their reservoir potential.
and Argentina. Three depositional successions (Silurian– Recent outcrop studies along the margins of the basin
Devonian, Upper Carboniferous–Jurassic, and Creta- have identified large glaciofluvial feeder channels, the
ceous) record repeated phases of subsidence and locations of which are controlled by basement structures
sediment accommodation in the Paraná basin (Oliveira, (Figure 10)‚ but a channeled geometry has so far not been
172 Eyles et al.

Figure 6—Paraná basin of Brazil showing


location of wells used in this study and
representative stratigraphic and resistivity
logs through the Itararé Group in the central
part of the Paraná basin at well 3. Inset
shows the location of Figure 10 in the state
of Mato Grosso.

recognized in the subsurface given the dearth of high- suspended sediment plumes; laminated facies were
quality seismic coverage. Initial interpretations of the deposited from turbidity currents. Shales in general have
thick sandstones of the Itararé Group indicate deposition low total organic carbon content, but form excellent seal
within large submarine channels and associated lobe rocks.
systems controlled by reactivated basement structures
(see Figure 11). Depositional Setting
Diamictite beds consist of scattered clasts in a clayey Primary glacial deposits such as tillites, boulder
silt to muddy sand matrix. Massive facies were pavements, and ice proximal conglomerates and sand-
deposited by the rain-out of mud from suspended stones, together with erosional bedforms such as
sediment plumes together with ice-rafted debris. Strati- glacially scoured basement highs, are well exposed
fied and locally laminated facies record postdepositional around the main outcrop belt of the eastern Paraná basin
downslope redeposition as debris flows. These form between Curitiba and São Paulo (see Rocha-Campos and
blanket-like deposits in the Itararé Group and consist of dos Santos, 1981; dos Santos, 1987). It is probable that this
millimeter- to boulder-sized fragments set in an argilla- zone of ice contact deposits and landforms marks the
ceous matrix. Clast content varies widely; diamictites westward limit of ice flowing out of southern Africa
pass laterally and vertically into marine shales. In many during basin filling (Kaokoveld ice lobe of Franca and
cases, discrimination of rain-out and resedimented Potter, 1991). However, as recognized in southern Africa
components in many diamictite units is not possible and by Visser (1989), basin filling may have followed a long
the term rain-out complex is used. Similar stratigraphic period (20 Ma) of extensive glaciation, uplift, and erosion
complexes composed of in situ rain-out and slumped across the Paraná and Karoo basins for which no sedi-
facies dominate the correlative Dwyka Formation of mentary record survives.
southern Africa (Visser, 1989) and are also reported from During filling, the Paraná basin acted as a steep-sided
many other glaciated basins (e.g., Young and Gostin, trap for prolific volumes of glacially produced sediment,
1991; Eyles, 1993). Typical gamma ray, resistivity, and primarily sand and mud, delivered by glacial meltwaters
dipmeter logs of diamictites are shown in Figure 8c. (Figure 11). The alternation of sandstone and shale
Dipmeter logs are patternless, probably as a result of members in each formation most likely records an
extensive resedimentation, whereas other geophysical overall long-term tectonic control rather than any alter-
logs show a fairly uniform pattern reflecting a fine- nation of climate (glacial–interglacial cycles) or climati-
grained, shale-like matrix. cally driven glacioeustatic or glacioisostatic changes.
Shales form massive blanket-like units across the Stratigraphic cross sections show that shales overstep
basin and accumulated by the settling of fines from sandstones (see Eyles et al., 1993), which indicates basin
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 173

Figure 7—Representative sedi-


mentologic logs of cores
through diamictites (seals) and
turbidite sandstones (reser-
voirs) of the Itararé Group. (a)
Well #2, Cahpeu do Sol
Member. (b) Well #2, Rio
Segredo Member. (c) Well #1,
Cuiba Paulista Member. See
Figure 6 for well locations.
Lithofacies codes are after
Eyles et al. (1983); numbers on
left are depths in meters. Strati-
graphic position of log (c) is
shown on Figure 8a.

subsidence and transgression of the coastline. Short-term Chaco-Paraná Basin of Argentina


climate and sea level changes are beyond the resolving
power of the large-scale basinwide work conducted to The Chaco-Paraná basin contains at least a 2.2-km
date, but given the relatively steep margins of the Paraná thickness of Late Carboniferous–Late(?) Permian strata
basin, the sedimentary effects of such sea level changes present in the subsurface of the Chaco-Paraná plains in
are probably slight. This situation can be contrasted with central and northern Argentina (Figures 1, 12a). Basin
that obtained in the foreland basins of Argentina, where structure and stratigraphy are poorly known from
a high-resolution sea level and climate record may ulti- wildcat drilling and low-density seismic coverage. Mild
mately be resolvable from shallow water, glacially influ- flexural deformation and tilted fault blocks can be seen in
enced shelf strata. Drawing on a wide range of data, seismic sections (Pezzi and Mozetic, 1989). The thickest
Eyles (1993) reviewed the available database and measured section is 2200 m in the Ordoñez stratigraphic
suggested that the maximum glacioeustatic sea level well (Figure 12b).
drawdown resulting from late Paleozoic glaciation was Late Paleozoic sections in wells in the northern Chaco-
unlikely to have been much greater than 70 m. Paraná basin are subdivided into three conformable
174 Eyles et al.

(a) (b)

(c)

Figure 8—Stratigraphic and geophysical logs. (a) Sand-


stones of the Cuiaba Paulista Member, well #1. Black bar
indicates cored interval shown in detail in Figure 7c. Note
distinct cylindrical gamma ray pattern between shale
intervals and consistent dipmeter values. Sandstones are
fine to medium grained and texturally mature and record
repeated stacking of Bouma A turbidites, which have
excellent reservoir potential. (b) Campo Mourao Member,
well #5. Fining-upward (FU) cycles identify thick, graded
sandstone units; these are less texturally mature than
those in part (a) and are less attractive exploration targets.
Note greater scatter of dipmeter values compared to part
(a). (c) Diamictites of the Chapeu do Sol Member, well #5.
Note wider scatter of dipmeter values reflecting wide-
spread debris flow during deposition of rain-out diamictite
complexes and shale-like gamma ray log pattern reflecting
fine-grained muddy matrix. See Figure 6 for well locations.

formations (from the bottom up): Sachayoj, Charata, and Chaco-Tarija Basin of Bolivia and
Chacabuco. In contrast, correlative strata in the south are Argentina
grouped into the Ordoñez and the conformably overlying
Victoriano and Rodriguez formations (Mingramm et al., The Chaco-Tarija basin is an asymmetric intracratonic
1979; Russo et al., 1979) (Figures 2, 12b). Well logs show basin that extends more than 1000 km from Peru in the
alternating packets of mudstone, sandstone, and diamic- north to Argentina in the south (Figures 1, 13, 14). Figure
tite containing marine faunas. Stratigraphic, sedimento- 1 shows a much simplified assessment of the structural
logic, and structural similarities with the Paraná basin fill setting of the basin emphasizing the first-order control by
suggest that the same depositional systems developed in late Proterozoic terranes (see Tankard et al., 1995). The
both basins. Nevertheless, both basins appear to have northern boundary of the Chaco is defined by the linear,
been separated by the Asunción arch and likely reflect lineament-defined margin of the Guapore shield of
subsidence of independent fault-bounded subbasins Brazil; to the south, the basin is closed by the Puna and
controlled by late Proterozoic lineaments (F. Wiens, Michicola arches in northwestern Argentina.
personal communication, 1992). In the Chaco-Traija basin, strata deformed during the
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 175

Figure 9—Secondary moldic porosity resulting from


dissolution of feldspar grains in reservoir sandstones of
the Rio Segredo Member of the Itararé Group. Molds are
identifiable by remnant rims of feldspar overgrowth which
are more stable than original detrital grains. P, porosity; Q,
quartz; F, feldspar. Scale bar in upper left is 0.1 mm long.

Chanica orogeny are overlain by a thick Carboniferous


sequence, which commenced deposition in the Early
Carboniferous (Salfity et al., 1987; Suarez, 1989) (Figure 2).
Carboniferous strata are well exposed along the sub-
Andean foothills. Outcrops show muddy diamictite
facies and large sandstone-filled channels up to several
tens of kilometers wide and several hundred kilometers
long (Figure 13), commonly showing large-scale defor-
mation structures and chaotic bedding. In the past, these Figure 10—Generalized outcrop map near Cuiaba, Mato
channels have been interpreted as continental tillites and Grosso, along the northwestern margin of Paraná basin
subaerial meltwater channels, respectively, deposited by (see inset in Figure 6 for location) showing basement
ice margins advancing and retreating across the Chaco structural control on location of sandstone-filled feeder
basin from ice centers to the south and east (Helwig, channels (see Figure 11). (From N. Eyles, unpublished
1972; Salinas et al., 1978). New work (discussed below) data, 1991.)
stresses instead the importance of marine sedimentation
and the paucity of evidence for any direct glacial
dence. Slope failure along the frontal margins of deltas
influence on sedimentation.
may have been the principal mechanism for generating
thick, sandy turbidites. Seismic lines and outcrops show
Facies and Depositional Setting chaotically deformed masses (olistostromes) resulting
Generalized stratigraphic columns based on outcrops from downslope collapse of channel walls or the upslope
in the sub-Andean foothills near Santa Cruz, Bolivia basin margin, most likely in response to large-magnitude
(Figure 15), show thick, channeled sandstones (e.g., earthquakes.
Itacua and Chorro formations) separated by fine-grained Fine-grained intervals that separate sandstones (e.g.,
facies including pebbly mudstones and shales (e.g., Tarija Taiguati Formation) (Figure 15) appear to have been
and Taiguati formations). The Tupambi and Tarija sand- deposited in interchannel settings. These facies
stones form prominent escarpments along the foothills commonly contain rafts and pillows of displaced
belt; shales form recessive units along valley floors. sandstone and muddy diamictite. The latter may record
Sandstones are dominated by thick, amalgamated the “spilling” of coarse-grained turbidites from large
sections of turbidite facies commonly deformed by channels which, as a consequence of being rapidly
slumping. Conglomerates are a minor component. dumped on a muddy substrate, underwent downslope
Amalgamated turbidite sandstones fill the large channels mass flow and mixing with mud. The generation of
discussed in the previous section (Figures 13, 16). The pebbly muds in this fashion is well known (e.g., Crowell,
regional depositional setting appears to have consisted of 1957; Eyles and Eyles, 1989; Eyles, 1990) and has previ-
sand-rich deltas around the basin margin in Argentina ously been suggested for the diamictites of the Chaco
funneling large volumes of sediment to a system of basin (Frakes and Crowell, 1969).
channels along the basin axis (Figure 17a). The great The same basic stratigraphy of large feeder channels
thickness of individual channel fills suggests repeated separated by muddy mass flow deposits can be recog-
stacking and amalgamation of turbidites and rapid subsi- nized in the southern part of the basin in Argentina
176 Eyles et al.

Figure 11—From glacial source to reservoir rock: depositional model for glacially influenced sandstone turbidites of the late
Paleozoic Itararé Group of Brazil. The key to deposition of thick sandstone members (Figures 7, 8) is the occurrence of a
major influx of glaciofluvial sediment and repeated downslope resedimentation by slumps and slump-triggered turbidites.
Position of major feeder channels is controlled by basement faults (Figure 10). Correlative channels are preserved along the
Namibian coastline of southern Africa. Slumps may be generated by earthquakes recording intracratonic faulting during
basin subsidence. Source rocks are Devonian and Permian shales; areally extensive blankets of shales and diamictites (not
shown) form seals. Alternation of sandstone and muddy members in the Itararé Group (Figure 6) may reflect low relative sea
level and progradation of braid deltas, followed by highstands of relative sea level and deposition of transgressive muds.

(Lopez Gamundi, 1986; Starck et al., 1992). There, the Rocha-Campos et al. (1977) described marine fossils,
base of the succession (Tupambi Formation) contains including Levipustula levis Maxwell and Limipecten cf. L.
shallow water indicators and striated boulder pavements; burnettensis Maxwell, from the Taiguati and San Telmo
available data suggest a large deltaic system in close formations (Figure 16) (see also Sempere, 1995). These
proximity to ice and meltwaters. It is likely that this area authors summarize previous work, arguing for a marine
represents the source area for the channeled sandstones connection to the Chaco-Tarija basin and discuss a range
deposited to the north in Bolivia (Figures 13, 17a). of paleogeographic settings. The considerable thickness
It is not yet possible to reconstruct the larger deposi- (hundreds of meters) of sandstone-filled channels in the
tional system or water depths within the Chaco-Tarija Chaco-Traija basin suggests rapid subsidence. Deposi-
basin. The orientation and distribution of channels in the tion occurred within an extensional tectonic regime
basin (Figure 13) suggests that the term fan is not appro- involving rapid basement subsidence in response to
priate because the characteristic diverging pattern of transpressional movements along the nearby Pacific
distributary channels and a single source typical of plate margin. A marine connection to the northwest to
classic fan bodies cannot be identified. “Classic” fans the paleo-Pacific Ocean appears likely, but the basin
develop on simple, unconfined basin floors with a single proper may have had a restricted (brackish) circulation
feeder channel (e.g., Walker, 1992) in contrast to other system dominated by freshwater input from terrestrial
fans that occupy confined basins and do not show a ice centers to large deltas (Figure 17a).
simple diverging network of channels.
A useful morphologic analog for the channel system of
the Chaco-Tarija basin is provided by the eastern
Canadian continental slope and basin plain, which has HYDROCARBONS AND RESERVOIR
experienced repeated Pleistocene glaciation. Hesse and CHARACTERISTICS
Rakofsky (1992) have described a basin-wide network of
converging and subparallel (“Yazoo type”) submarine Foreland basins
channels cut on the floor of the Labrador Sea (Figure
17b). Thick, massive turbidite sands comprise the In the foreland basins of Argentina, stratigraphically
dominant in-channel facies and interfinger with chaoti- complex glacial deposits are poorly prospective because
cally bedded slump facies resulting from the collapse of of a dearth of reservoir quality sandstones and appro-
channel walls. Fine-grained, muddy turbidite facies priate source and seal rocks. In addition, as a conse-
comprise typical overbank facies. Inactive channels are quence of Andean tectonics, most potential reservoirs
buried below thick, mud-rich debris flow deposits origi- have been breached by erosion. Secondary high-risk
nating from the downslope collapse of glacial sediments targets for exploration probably occur in the Andean
dumped along the upper continental slope. These facies foredeep and in footwall basins where upper Paleozoic
are comparable to those of the Chaco-Tarija basin. sandstones lie adjacent to Triassic source shales. Only in
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 177

Figure 12—(a) Chaco-Paraná


basin of Argentina showing
isopachs for upper Paleozoic
fill and principal structural
elements that controlled sedi-
mentation. Open circles
indicate position of exploration
wells. Solid triangles show
presence of diamictite in
subsurface. (b) Stratigraphy for
southern Chaco-Paraná basin
at a YPFB exploration well.
(After Russo et al., 1980.)

the southern Tepuel basin is there some potential for


exploration in postglacial, shallow marine sandstones
winnowed from underlying glacial strata. Porosity is
low, however, ranging from 3 to 8% due to cementation
by silica and micaceous minerals.

Intracratonic Basins
In the intracratonic basins of Argentina, Brazil, and
Bolivia, the principal role of late Paleozoic glaciers has
been to scour surrounding highlands and cratons and to
furnish large volumes of glaciofluvial sand and mud to
rivers feeding coastal deltas and their channelized
submarine equivalents (Figures 11, 17a). Repeated
sediment instability was caused by faulting (earth-
quakes?) and downslope reworking of glaciofluvial
sediment by sediment gravity flows along steep fault-
controlled basin margins. These processes resulted in the
focusing of sand along structural lineaments and the
formation of thick, prospective sandstone bodies.

Chaco-Tarija Basin
In the Chaco-Tarija basin, the Bolivian National Oil
Company (Yacimientos Petroliferos Fiscales Bolivianos,
or YPFB) recovers hydrocarbons from glacially influ-
enced submarine channel fills (Figures 13, 16, 17a). These
Figure 13—Distribution of Carboniferous strata in the strata host significant reservoirs in about 30 oil fields
Chaco basin of Bolivia and schematic distribution of large- (e.g., McCaslin, 1979; Montes de Oca, 1989) and are
scale submarine channels. (After Salinas et al., 1978; sourced from Devonian shales. Oil fields are located in
Tankard et al., 1995. See also Eyles, 1993.) thrust sheets with diamictites as seals. Westward tilting
178 Eyles et al.

Figure 15—Representative stratigraphic columns for the


late Paleozoic in the Argentinian (after Starck et al., 1992)
and Bolivian sectors (Santa Cruz district; after López
Paulsen et al., 1992) of the Chaco-Tarija basin. Note
differing use of formation names in Bolivia and Argentina.
Devonian shales are source rocks (see Figure 14b). The
Tupambi Formation is productive in the Campo Durán-
Madrejones oil field in Argentina. Note coarsening-upward
Figure 14—(a) Chaco-Tarija basin of Argentina and Bolivia
trend from the Itacua to Escarpment Formation probably
showing isopachs for Carboniferous fill and principal
recording progradation of shoreline. Marine fossils are
structural elements that controlled sedimentation. The
reported from the Taiguati and San Telmos formations
basin boundary in the northwest is ill defined. Horizontal
(Rocha-Campos et al., 1977).
ruling shows distribution of thermally mature Devonian
seals. Black ovals show approximate extent of principal oil
and gas fields in Carboniferous. (b) Thermal gradients in turbidite sandstones within the Cuiaba Paulista, Campo
Chaco-Tarija basin. Present oil window (65–100˚C) includes Mourao, and Rio Segredo members of the Itararé Group.
shaly Devonian sections. Position of reconstructed Late In general, reservoir quality is fair to poor because of sili-
Cretaceous 100˚C isotherm suggests maturation
cification and mechanical compaction. Along the central
throughout the Cenozoic. (After Salinas et al., 1978.)
parts of the basin, sandstones are present at depths as
great as 4600 m. Thick, texturally mature sandstone
of the basin during late Cenozoic Andean deformation bodies having cylindrical gamma ray logs (Figure 8a) are
led to migration of hydrocarbons to higher stratigraphic the best targets; other sandstones showing well-defined
levels within thrust sheets. Thermal conditions in the fining-upward characteristics (Figure 8b) offer less attrac-
basin throughout the Cenozoic have been favorable for tive targets because of their higher matrix content.
maturation of Devonian source rocks (Figure 14b). The Campo Mourao Formation (Figure 6) contains the
Because of repeated cutting and filling and the evolution majority of potential reservoirs. This is related to the
of multistory channels fills, stratigraphic correlations development of secondary porosity caused by dissolu-
from sub-Andean outcrops to seismic sections within the tion of early calcite and anhydrite cements and, to a
basin are difficult. Most oil finds were made in the course lesser extent, dissolution of feldspars and unstable lithics
of drilling to deeper Devonian targets. during intrusion of Jurassic–Cretaceous dikes and sills
(Figure 9). Large volumes of carboxylic acids and
Paraná Basin groundwater rich in carbon dioxide were released from
In Brazil, more than 110 exploration holes have been the intrusives as a result of the maturation of organic
completed to date in the Paraná basin, with the most matter (França and Potter, 1991). The best reservoirs of
important reservoir targets being the thick, amalgamated the Itararé Group occur in the uppermost sandstones of
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 179

plots show that the Ponta Grossa Formation entered the


oil window at about 200 Ma and remained there until
about 135 Ma; the Irati Shale entered the oil generation
window at about this time (Oliveira, 1989). Intrusion of
the Serra Geral lavas played a major role in elevating
maturation levels by direct thermal heating and by accel-
erating basin subsidence. The base of the Itararé Group is
beyond the preservation limit for wet gas, but several
holes show dry gas composed largely of methane
(typically about 85%), with less than 5% of ethane,
propane, isobutane, and normal butane. The most
(a) promising test to date shows a gas flow rate of 51,000
m3/day in sandstones of the Campo Mourao Formation,
where a diabase sill forms a seal (Zalán et al., 1990).
Exploration in the Paraná basin is complicated by the
thick cover of Upper Jurassic Serra Geral lavas, but
underlying upper Proterozoic basement structures
appear to have dictated the geometry of successive
phases of late Paleozoic extension and basin expansion
(Eyles and Eyles, 1993). Oliveira (1989) presented the
results of backstripping analyses and showed that the
greatest amount of subsidence between the Silurian and
the Cretaceous occurred during deposition of the
Permian–Carboniferous Itararé Group. Hercynian oroge-
nesis along the paleo-Pacific margin of southern South
(b) America resulted in the “far-field” reactivation of the
Curibita-Guapiara structural zone and associated linea-
ments (Figure 3). The changing state of stress within the
interior of the continent as Gondwana underwent a
clockwise rotation may have controlled renewed
intracratonic subsidence.
In the Paraná basin, thickness isopachs for successive
stratigraphic members show a strong basement control
on both depocenters and sediment sources during depo-
sition of the Itararé Group (see Eyles et al., 1993, for
details). The models of Prior and Bornhold (1986) and
Syvitski and Farrow (1989) describe strongly focused
sedimentation dominated by mass flow within the
(c) confines of steep-sided glaciated fiord basins. These
models may be appropriate for sandstones of the Itararé
Figure 16—(a) Base of turbidite-filled channel cutting Group in the Paraná basin (Figure 11). The same reacti-
across slumped sandstones and diamictites (Tarija vation processes on a different scale have probably
Formation, near Samaipata, Bolivia). Section is 30 m high. controlled fluid migration along structural lineaments.
(b) Olistostromes of turbidite sandstone and diamictite, Faulting also results in abrupt juxtaposition of sandstone
Tarija Formation. Section is 15 m high. (c) Pebbly against fine-grained facies.
mudstone debris flow of the Tupambi Formation with bed
A clear relationship among basement lineaments,
base identified by clasts and interbedded graded (turbidite)
sandstones. paleocurrents, and sandstone body geometry is evident
in outcrops along the northwestern outcrop belt of the
Itararé Group near Rondonopolis (Figure 10). The
the Rio Segredo Member (Figure 6). Because of their positions of large sandstone-filled feeder channels are
shallow depth, these show the least mechanical controlled by grabens in the underlying basement
compaction of all the Itararé sandstones and also have (Figure 11). Similar structural relationships between
little clay matrix. The cleanliness of the sandstones is the basement and so-called preglacial valleys crop out along
result of elutriation of fines by repeated downslope the southeastern basin margin south of São Paulo (e.g.,
turbidity flow. Rio Segredo sandstones are sealed by Martin, 1953, 1964; Martin et al., 1958; Mau, 1960). These
thick (10 m average), extensive (>600 km2) diamictites. preserve coarse-grained, subaqueous mass flow and
Devonian shales of the Ponta Grossa Formation are subaerial outwash facies and, most significantly, can be
the principal source rocks for gas and condensate shows correlated with other paleovalleys exposed in north-
in the Itararé Group (Zalán et al., 1990). Local oil finds in western Damaraland, Namibia (Horsthemke et al., 1990).
the overlying postglacial Rio Bonito Formation are Valleys are structurally controlled by northwest-trending
sourced from the Upper Permian Irati Shale. Geohistory late Proterozoic basement lineaments parallel to the
180 Eyles et al.

(a) (b)

Figure 17—(a) Generalized depositional setting for the


glacially influenced strata of the Chaco-Tarija basin. (b)
Yazoo-type channels on the floor of the Labrador Sea (after
Hesse and Rakofsky, 1992). Compare with channel
systems shown in Figure 13. The tectonic and bathymetric
setting is different from the Chaco-Tarija basin and only the
gross disposition of channels is compared here.

Guapiara lineament in Brazil and the trend of the cratons, thereby focusing glacial meltwaters and
Damara belt in southern Africa (see Tankard et al., 1995, sediments into structurally controlled depocenters. Other
and figure 16.9 in Eyles, 1993). In the Paraná basin, they basins lacking any basement control on drainage appear
contain significant postglacial coal deposits. to lack the means of focusing sediment and water,
With only about 30 wildcat wells, the Chaco-Paraná resulting in dispersal of coarse sediment around the
basin of Argentina is the least studied of all the glacially basin margins. Structurally controlled sources and
influenced basins in southern South America. As a result depocenters appear to be persistent through time and
of lithostratigraphic similarities and lack of independent may promote basin subsidence as a result of rapid
detailed information, exploration targets and inferred sediment loading adjacent to faulted basement. In turn,
reservoir charactistics are likely to be similar to those this may enhance sediment instability and downslope
described for the Paraná basin of Brazil. mass flow by earthquakes, leading to the accumulation
of thick turbidite sandstones. The axiomatic assumption
made by many petroleum geologists that the terms
DISCUSSION glaciated basin and poorly prospective are synonomous is
clearly misplaced.
The tectonostratigraphic model established here for
prospective glacigenic sandstones in intracratonic basins
in central and southern South America may be of wider
application. Broadly similar tectonic controls on the
formation of glaciated intracratonic basins in Australia Acknowledgments N. Eyles and C. H. Eyles thank the
can be identified (Eyles, 1993), and a common character- Natural Science and Engineering Research Council of Canada
istic of several of these basins is the presence of thick, for funding fieldwork in South America and for providing an
prospective sandstones (e.g., Grant Group of the International Scientific Exchange award to support G.
Canning basin; O’Brien, 1992). In contrast, other late González Bonorino’s sabbatical leave at the University of
Paleozoic glacially influenced basins are dominated by Toronto. González Bonorino thanks NSERC for making this
muddy facies, such as shales and rain-out diamictite collaborative work possible and the Eyles for their hospitality.
complexes, with an absence of thick sandstones (e.g., França’s stay at Toronto was funded by Conselho Nacional de
Karoo basin; Visser, 1989). This contrast can be argued to Desenvolvimento Cientifico e Tecnologo, Petrobras, and
be a direct result of regional basement structure and the NSERC. We thank Tony Tankard, Edison Milani, Paul Potter,
presence or absence of major lineament systems that Paulo dos Santos, Tony Rocha-Campos, and Fernando Wiens
controlled drainage patterns, sediment sources, and for discussions. The ideas presented herein are those of the
depocenters. Eyles et al. (1993) showed that basement authors. The manuscript was reviewed by Johan Visser, Tony
lineaments play a fundamental role in basin filling by Tankard, and Barend van Hoorn, whom we thank for their
capturing meltwater runoff from glacially scoured valuable comments.
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 181

REFERENCES CITED setting: Palaeogeography, Palaeoclimatology, Palaeoe-


cology, v. 79, p. 73–98.
Andreis, R. R., Archangelsky, S., González, C. R., López Eyles, N., 1993, Earth’s glacial record and its tectonic setting:
Gamundi, O., and N. Sabattini, 1987, Cuenca Tepuel- Earth Science Reviews, v. 35, p. 1–248.
Genoa, in S. Archangelsky, ed., El sistema Carbonifero en Eyles, N., and C. H. Eyles, 1989, Glacially influenced deep
la República Argentina: Academia Nacional de Ciencias, marine sedimentation of the late Precambrian Gaskiers
Córdoba, p. 166–196. Formation, Newfoundland, Canada: Sedimentology, v. 36,
Archangelsky, S., C. L. Azcuy, C. R. González, and N. p. 601–620.
Sabattini, 1987, Correlacion general de biozonas, in S. Eyles, N., and C. H. Eyles, 1992, Glacial depositional systems,
Archangelsky, ed., El sistema Carbonifero en la República in R. G.Walker and N. P. James, eds., Facies models:
Argentina: Academia Nacional de Ciencias, Córdoba, response to sea-level change: Geological Association of
p. 281–292. Canada Special Publication, p. 73–100.
Azcuy, C. L., R. R. Andreis, A. Cuerda, M. A. Hunicken, M. V. Eyles, N., and C. H. Eyles, 1993, Glacial geologic confirmation
Pensa, D. A. Valencio, and J. F. Vilas, 1987, Cuenca of an intraplate boundary crossing the Paraná basin of
Paganzo, in S. Archangelsky, ed., El sistema Carbonifero Brazil: Geology, v. 21, p. 459–462.
en la República Argentina: Academia Nacional de Eyles, N., C. H. Eyles, and A. D. Miall, 1983, Lithofacies types
Ciencias, Córdoba, p. 41–99. and vertical profile modes: an alternative approach to the
Bossi, G., and R. R. Andreis, 1985, Secuencias delteaicas y description and environmental interpretation of glacial
lacustres del Carbonifero del centro-oeste Argentino, X: diamict sequences: Sedimentology, v. 30, p. 393–410.
Congreso International de Estratigrafia y Geologia del Frakes, L. A., and J. C. Crowell, 1969, Late Paleozoic glacia-
Carbonifero, Madrid, v. 3, p. 285–309. tion: I, South America: GSA Bulletin, v. 80, p. 1007–1042.
Brodzikowski, K. and A. J. Van Loon, 1991, Glacigenic Frakes, L. A., and J. C. Crowell, 1972, Late Paleozoic glacial
sediments, in Developments in Sedimentology, Elsevier, geography between the Paraná basin and the Andean
v. 49, 674 p. geosyncline: Anales Academie Brasileira Ciencias, v. 44,
Caputo, M. V., and J. C. Crowell, 1985, Migration of glacier p. 139–145.
centers across Gondwana during Paleozoic era: GSA França, A. B., and P. E. Potter, 1991, Stratigraphy and
Bulletin, v. 96, p. 1020–1036. reservoir potential of glacial deposits of the Itararé Group
Coleman, A. P., 1926, Ice Ages: Recent and Ancient: New (Carboniferous–Permian), Paraná basin, Brazil: AAPG
York, MacMillan, 296 p. Bulletin, v. 75, p. 62–85.
Crowell, J. C., 1957, Origin of pebbly mudstone: GSA Bulletin, Goldstein, B. A., 1989, Waxing and waning in stratigraphy,
v. 68, p. 993–1010. play concepts and prospectivity in the Canning basin:
Crowell, J. C., 1983, Ice ages recorded on Gondwanan conti- Australian Petroleum Exploration Association Journal, v.
nents: Transactions of the Geological Society of South 29, p. 466–508.
Africa, v. 86, p. 237–262. González, C. R., 1981, Pavimento glaciario en el Carbónico de
Daemon, R. F., L. P. Quadros, A. T. Picarelli, M. M. Toigo, and la Precordillera: Revista de la Asociación Geológica
M. C. Klepzig, 1992, Arcabouco Bioestratigrafico da Bacia Argentina, v. 36, p. 262–266.
do Paraná-Grupo Itararé: Petrobras Report Projeto 01. 02. González, C. R, 1990, Development of the late Paleozoic
42, 110 p. glaciations of the South American Gondwana in western
Derby, O. A., 1888, Spuren einer Carbonen eiszeit in Argentina: Palaeogeography, Palaeoclimatology, Palaeoe-
Sudamerika: Neues Jahrbuch fur Min., v. 2, p. 172–176. cology, v. 79, p. 275–287.
Diaz, E., and J. C. Lema, 1991, Diamictitas Glaciomarinas en el González Bonorino, F., and G. González Bonorino, 1988, La
Carbonifero del Altiplano norte de Bolivia: sedimentología base del Grupo Tepuel en las cercanias de Esquel, Chubut:
e interpretación de ambientes sedimentarios, 6to: Congreso Revista de la Asociación Geológica Argentina, v. 43,
Geológico Chileno, Viña del Mar. p. 518–528.
Dickins, J. M., 1985, Late Paleozoic glaciation: Bureau of González Bonorino, G., 1991, Late Paleozoic orogeny in the
Mineral Resources Journal of Australian Geology and northwestern Gondwana continental margin, western
Geophysics, v. 9, p. 163–169. Argentina and Chile: Journal of South American Earth
Dickins, J. M., 1993, Climate of the Late Devonian to Triassic: Sciences, v. 4, p. 131–144.
Palaeogeography, Palaeoclimatology, Palaeoecology, González Bonorino, G., 1992, Carboniferous glaciation in
v. 100, 89–97. Gondwana: evidence for grounded marine ice and conti-
dos Santos, P. R., 1987, Facies e evolucao paleogeografica do nental glaciation in southwestern Argentina: Palaeogeog-
subgrupo Itararé/grupo aquidauana (Neopaleozoico) na raphy, Palaeoclimatology, Palaeoecology, v. 91, p. 363–375.
Bacia do Paraná, Brasil: Ph.D. dissertation, University of González Bonorino, G., G. Rafine, V. Vega, and D. Guerin,
São Paulo, São Paulo, Brazil, two volumes. 1988, Ambientes de plataforma neritica dominada por
Du Toit, A. L., 1927, A geological comparison of South tormentas en la sección glacigénica del Grupo Tepuel
America with South Africa: Carnegie Institution, Publica- (Paleozoico superior), en las sierras de Tepuel y de Tecka,
tion 381, Washington, D.C., 128 p. Chubut noroccidental, Argentina: Revista de la Asociación
Eyles, C. H., N. Eyles, and A. D. Miall, 1985, Models of Geológica Argentina, v. 43, p. 239–252.
glaciomarine deposition and their applications to ancient Grahn, Y., and Caputo, M. V., 1992, Early Silurian glaciations
glacial sequences: Palaeogeography, Palaeoclimatology, in Brazil: Palaeogeography, Palaeoclimatology, Palaeoe-
Palaeoecology, v. 51, p. 15–84. cology, v. 99, p. 9–15.
Eyles, C. H., N. Eyles, and A. B. França, 1993, Late Paleozoic Hambrey, M. J., and W. B. Harland, eds., 1981, Earth’s Pre-
glaciation in an active intracratonic basin—Itararé Group, Pleistocene Glacial Record: Cambridge, Cambridge
Paraná basin, Brazil: Sedimentology, v. 40, p. 1–25. University Press, 1004 p.
Eyles, N., 1990, Late Precambrian “tillites” of the Avalonian- Helwig, J., 1972, Stratigraphy, sedimentation, paleogeog-
Cadomian belt; marine debris flows in an active tectonic raphy, and paleoclimates of Carboniferous (“Gondwana”)
182 Eyles et al.

and Permian of Bolivia: AAPG Bulletin, v. 56, O’Brien, P. E., J. F. Lindsay, P. N. Southgate, M. J. Jackson, J.
p. 1008–1033. M. Rennard, and M. J. Sexton, 1992, Sequence stratigraphy
Hesse, R., and A. Rakofsky, 1992, Deep-sea of glacial sediments in an intracratonic basin—Grant
channel/submarine Yazoo system of the Labrador Sea: A Group, Canning basin, Western Australia: AAPG Bulletin,
new deep-water facies model: AAPG Bulletin, v. 76, v. 76, p. 1120.
p. 680–707. Oliveira, L. O. A., 1989, Aspectos da evoluçao termomecanica
Horsthemke, E., S. Ledendecker, and H. Porada, 1990, Deposi- da bacia do Paraná no Brasil: Revista Brasileira de Geocien-
tional environments and stratigraphic correlation of the cias, v. 19, p. 330–342.
Karoo Sequence in northwestern Damaraland: Communi- Page, R. F. N., C. O. Limarino, O. López Gamundi, and S.
cations of the Geological Survey of Namibia, v. 6, p. 63–73. Page, 1984, Estratigrafia del Grupo Tepuel en su perfil tipo
Langford, R. P., 1991, Permian coal and paleogeography of y en la región El Molle, provincia del Chubot: IX Congr.
Gondwana: Australian Bureau of Mineral Resources Geol. Argentino, Bariloche, v. I, p. 619–632.
Record, 1991/95, p. 136. Pezzi, E. E., and M. E. Mozetic, 1989, Cuencas sedimentarias
Leighton, M. W., and D. R. Kolata, 1990, Selected interior de la región Chaco, in G. Chebli and L. A. Spalletti, eds.,
cratonic basins and their place in the scheme of global Cuencas Sedimentarias Argentinas: Universidade
tectonics: a synthesis, in M. W. Leighton, D. R. Kolata, D. F. Nacional de Tucumán, Argentina, p. 65–78.
Oltz, and J. J. Eidel, eds., Interior cratonic basins: AAPG Prior, D. B., and B. D. Bornhold, 1986, Sediment transport on
Memoir 51, 729–799. subaqueous fan delta slopes, Britannia Beach, British
Levell, B. K., J. H. Braakman, and K. W. Rutten, 1988, Oil- Columbia: Geo-Marine Letters, v. 5, p. 217–224.
bearing sediments of Gondwana glaciation in Oman: Ramos, V. A., 1989, The birth of southern South America:
AAPG Bulletin, v. 72, p. 775–796. American Scientist, v. 77, p. 444–450.
López Gamundi, O., 1983, Modelo de sedimentación glacima- Redfern, J., 1991, Subsurface facies analysis of Permo-
rina para la Formación Hoyada Verde, Paleozoico superior carboniferous glacigenic sediments, Canning basin,
de la provincia de San Juan: Revista de la Asociación Western Australia, in H. Ulbrich and A. C. Rocha-Campos,
Geológia Argentina, v. 38, p. 60–72. eds., Gondwana Seven Proceedings: Instituto de Geoscien-
López Gamundi, O., 1986, Sedimentologia de la Formación cias, Universidade de São Paulo, p. 349–363.
Tarija, carbonifero de la Siera de Aguarague, Provincia de Rocha-Campos, A. C., R. G. de Carvalho, and A. J. Amos,
la Salta, Argentina: Revista Asociación Geológia 1977, A Carboniferous (Gondwana) fauna from subandean
Argentina, v. 41, p. 334–355. Bolivia: Revista Brasileira de Geociencias, v. 7, p. 287–304.
López Gamundi, O., C. L. Azcuy, A. Cuerda, D. A. Valencio, Rocha-Campos, A. C., and P. R. dos Santos, 1981, The Itararé
and J. F. Vilas, 1987, Cuencas Río Blanco y Calingasta- Subgroup, Aquidauana Group, and San Gregorio
Uspallata, in S. Archangelsky, ed., El Sistema Carbonifero Formation, Paraná basin, southeastern South America, in
en la República Argentina: Academia Nacional de M. J. Hambrey and W. B. Harland, eds., Earth’s Pre-Pleis-
Ciencias, Córdoba, p. 101–132. tocene Glacial Record: Cambridge, Cambridge University
López Paulsen, O., M. López Pugliesi, R. Suárez Soruco, and J. Press, p. 842–852.
Oler Veramendi, 1992, Estratigrafía, facies, ambientes y Russo, A., R. Ferello, and G. Chebli, 1979, Llanura Chaco
tectónicas fan erozoicas en un sector de la Cordillera de los Pampeana, in Segundo Simposio de Geología Regional de
Andes, Sucre, Chuquisaca (abs.): I Conferencia Interna- Argentina: Academia Nacional de Ciencias, Córdoba, v. 1,
cional de las Cuencas Fanerozoicas del Gondwana sudoc- p. 139–183.
cidental, Santa Cruz, p. 56. Russo, A., S. Archangelsky, and J. C. Gamerro, 1980, Los
Martin, H., 1953, Notes on the Dwyka succession and some depósitos suprapaleozoicos en el subsuelo de la Ilanura
pre-Dwyka valleys in South West Africa: Transactions, Chaco-Pampeana, Argentina: 2° Congr. Argentino Paleon-
Geological Society of South Africa, v. LVI, p. 37–41. tologie, y Bioestratigraphie, v. 4, p. 157–173.
Martin, H., 1964, The directions of flow of the Itarare ice Salfity, J., C. L. Azcuy, O. López Gamundi, D. A. Valencio,
sheets in the Paraná basin: Boletin Paranaense de and J. F. Vilas, 1987, Cuenca Tarija, in S. Archangelsky, ed.,
Geografia, v. 12, p. 25–79. El Sistema Carbonifero en la Republica Argentina:
Martin, H., H. Mau, and A. J. S. Bjornberg, 1958, Vale pre- Academica Nacional de Ciencias, Córdoba, p. 15–39.
glacial a nordeste de Jundial, São Paulo: Boletin Societe Salinas, C., J. Oblitas, and C. Vargas, 1978, Exploración del
Brasilero Geologica, v. 8, p. 35–39. Sistema Carbonifero en la Cuenca Oriental de Bolivia:
Mau, H., 1960, Vale pre-glacial ao norte de Lavras do Sul, Rio Revista Técnica Yacimientos Petroliferos Fiscales Boli-
Grande do Sul: Boletin Societe Brasilero Geologica v. 9, vianos, v. 7, p. 5–49.
p. 79–82. Sempere, T., 1995, Phanerozoic evolution of Bolivia and
McCaslin, J. C., 1979, Eastern Bolivia’s Carboniferous system Adjacent regions, in A. J. Tankard, R. Suarez, and H. J.
draws interest: Oil and Gas Journal, v. 77, p. 199. Welsink, Petroleum basins of South America: AAPG
Milana, J. P., and F. Bercowski, 1987, Rasgos erosivos y Memoir 62, this volume.
depositacionales glaciales en el neopaleozoico de Slatt, R. M., 1984, Continental shelf topography: key to under-
Precordillera Central, San Juan, Argentina, in late standing distribution of shelf sand-ridge deposits from
Paleozoic of South America: IUGS-UNESCO Proy eto 211 Cretaceous Western Interior Seaway: AAPG Bulletin, v. 68,
Annual Meeting Abstracts, p. 56–59. p. 1107–1120.
Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979, Starck, D, E. Gallardo, and A. E. Schulz, 1992, La cuenca de
Sierras Subandinas, in Segundo Simposio de Geología Tarija: Estratigrafia de la proción Argentina: Boletin de
Regional de Argentina: Academia Nacional de Ciencias, informaciones Petroleras, Ano IX, no. 30, p. 2–14.
Cordoba, v. I, p. 95–137. Suarez, R., 1989, El ciclo cordillerano (Siluro-Carbonifero
Montes, de Oca, I. M., 1989, Geografia y. Recursos Naturales interior) en Bolivia y su relación con passes limitrofes:
de Bolivia: La Paz, Bolivia, Academia Nacional de Ciencias Revista Technica de Yaciementos Petrolíferos Fiscales
de Bolivia, 574 p. olivianos, v. 10, p. 233–243.
Hydrocarbon-Bearing Late Paleozoic Glaciated Basins of Southern and Central South America 183

Syvitski, J. P. M., and G. E. Farrow, 1989, Fjord sedimentation Authors’ Mailing Addresses
as an analogue for small hydrocarbon-bearing fan deltas,
in M. K. G. Whateley and K. T. Pickering, eds., Deltas: sites N. Eyles
and traps for fossil fuels: Geological Society of London,
Glaciated Basin Research Group
Special Publication 23, p. 21–43.
Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M. Department of Geology
Turic, A. B. Franca, E. J. Milani, B. B. de Brito Neves, N. Scarborough Campus
Eyles, H. de Santa Ana et al., 1995, Tectonic controls of University of Toronto
basin evolution in southwestern Gondwana, in A. J. Toronto, Ontario M1C 1A4
Tankard, R. Suarez, and H. J. Welsink, Petroleum basins of Canada
South America: AAPG Memoir 62, this volume.
Veevers, J. J., and C. Mc.A. Powell, 1987, Late Paleozoic glacial G. Gonzalez Bonorino
episodes in Gondwanaland reflected in transgressive-
Department of Geology
regressive depositional sequences in Euramerica: GSA
Bulletin, v. 98, p. 475–487. Conicet-University of Buenos Aires
Visser, J. N. J., 1989, Episodic Palaeozoic glaciation in the Buenos Aires 1428
Cape-Karoo basin, South Africa: in J. Oerlemans, ed., Argentina
Glacier Fluctuations and Climate Change: Boston, Kluwer,
p. 1–12. A. B. França
Visser, T., 1990, The age of the late Palaeozoic glacigene Petrobras/Nexpar
deposits in southern Africa: South African Journal of Rua Padre Camargo 285
Geology, v. 93, p. 361–375.
80060 Curitiba, PR
Walker, R. G., 1992, Turbidites and submarine fans, in R. G.
Walker and N. P. James, eds., Facies models—response to Brazil
sea level change: Geological Association of Canada Special
Publication, p. 239–264. C. H. Eyles
Wegener, A., 1929, The Origin of Continents and Oceans: Department of Geography
New York, Dover Publications, 260 p. McMaster University
Williams, B. P. J., E. K. Wild, and R. J. Suttill, 1985, Paraglacial Hamilton, Ontario L8S 4K1
aeolianites: potential new hydrocarbon reservoirs, Canada
Gidgealpa Group, southern Cooper basin: Australian
Petroleum Exploration Association Journal, v. 25,
p. 291–310. O. López Paulsen
Young, G. M., and V. A. Gostin, 1991, Late Proterozoic YPFB, Gerencia de Exploración
(Sturtian) succession of the North Flinders basin, South Santa Cruz
Australia: an example of temperate glaciation in an active Bolivia
rift setting, in J. B. Anderson and G. M. Ashley, eds.,
Glacial marine sedimentation: paleoclimatic significance:
GSA Special Paper 261, p. 207–223.
Youngs, B. C., 1975, The geology and hydrocarbon potential
of the Pedirka basin: Department of Mines, Geological
Survey of South Australia, Report of Investigations, v. 44,
p. 41.
Zalán, P. V., S. Wolff, M. A. M. Astolfi, I. S. Viera, J. C. J.
Concelçao, V. T. Appi, E. V. S. Neto, J. R. Cerqueria, and A.
Margues, 1990, The Paraná basin, Brazil, in M. W.
Leighton, D. R. Kolata, D. F. Oltz, and J. J. Eidel, eds.,
Interior cratonic basins: AAPG Memoir 51, p. 681–708.
Phanerozoic Tectonics and Sedimentation
in the Chaco Basin of Paraguay,
with Comments on Hydrocarbon Potential

F. Wiens
Geo Consultores
Asunción, Paraguay

Abstract

T his study of the Chaco basin is based on field studies of outcrops and on exploration data. The Chaco
basin covers 246,725 km2 of western Paraguay and consists of several depocenters or subbasins, each
with a unique tectonostratigraphic record. In the northwest, the Curupaity and Carandaity subbasins contain a
well-developed Paleozoic succession. In contrast, Mesozoic subsidence was marked in the southern Pirity and
Pilar subbasins and in the shallow Bahia Negra platform and San Pedro low to the east. These depocenters are
separated by structural highs. Uppermost Proterozoic–Recent sedimentary sequences are present in the Chaco
basin.
The subsidence history of the Chaco basin is recorded in four major unconformity-bounded sequences.
Northwest- and northeast-oriented structural lineaments of Eocambrian Brasiliano origin controlled the
patterns of subsidence. Mesozoic extensional tectonics related to the opening of the South Atlantic reorganized
the structural pattern of the Chaco basin; this episode is expressed in a system of half-grabens. Cenozoic
Andean orogenesis imposed the final structural readjustment and established the Chaco area as a modern
foreland basin.
Upper Devonian marine shales and Upper Cretaceous shales and carbonates are the primary source rocks
for hydrocarbons. The principal reservoir zones are Carboniferous channel sandstones in the Curupaity and
Carandaity subbasins and stratigraphic and structurally controlled sandstone reservoirs of Mesozoic age in the
Pirity subbasin.

Resumen

E ste trabajo se basa en datos de secuencias estratigraficas aflorantes y de actividades exploratorias. La


cuenca del Chaco ocupa con 246,725 km2 la región occidental del Paraguay y consiste de varios
depocentros o subcuencas, cada uno con un registro tectono-estratigráfico único. Al noroeste las subcuencas de
Curupaity y Carandaity representan áreas con secuencias paleozoicas bien desarrolladas. En contraste, subsi-
dencia mesozoica es marcada en las subcuencas de Pirity y Pilar al sur, como también en la plataforma de
Bahía Negra y el bajo de San Pedro al este. Estos depocentros están separados por altos estructurales. Sedi-
mentos desde el Proterozoico superior al reciente estan presentes en la cuenca del Chaco.
La historia de subsidencia de la cuenca del Chaco es registrada en cuatro ciclos secuencionales principales,
limitados por discordancias. Lineamentos estructurales orientados al noroeste y noreste, de origen eocámbrico
del ciclo Brasiliano, controlaron el estilo de la subsidencia. La tectónica distensional mesozoica relacionada a la
apertura del Atlántico Sur reorganiza la disposición estructural de la cuenca del Chaco. Este episodio es mani-
festado en un sistema de subcuencas asimétricas. La orogénesis Andina del Cenozoico impone los últimos
reajustes estructurales y establece el área del Chaco en una planicie promontoria moderna.
Lutitas marinas del Devónico superior y lutitas-carbonatos del Cretácico superior son las principales rocas
generadoras para hidrocarburos. Areas de mayor potencial de reservorio son areniscas de paleocauces
carboníferos en las subcuencas de Curupaity-Carandaity, y trampas estratigráficas-estructurales del Mesozoico
en la subcuenca de Pirity.

Wiens, F., 1995, Phanerozoic tectonics and sedimentation in the Chaco basin of Paraguay, 185
with comments on hydrocarbon potential, in A. J. Tankard, R. Suárez S., and H. J.
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 185–205.
186 Wiens

INTRODUCTION

The Chaco basin covers more than 60% of the


Republic of Paraguay, an area of 246,725 km2 (Figure 1).
The Chaco is a broad Quaternary plain with an average
elevation of only 160 m above sea level. Outcrops of
Paleozoic and Mesozoic strata occur in the northern
Chaco and along the Paraguay River (Figure 2). The
Paraguayan Chaco borders Bolivia to the north and west.
To the south and east, the Chaco continues into
Argentina and connects to eastern Paraguay and Brazil.
Hence, it is also known as the Chaco-Paraná basin.
Interpretation of the stratigraphic and tectonic
evolution is based on outcrop as well as subsurface
hydrocarbon and groundwater exploration data (PNUD,
1978; Wiens, 1989). Several hundred water wells have
been drilled throughout the Chaco. Hydrocarbon explo-
ration activities from 1947 to 1993 have contributed a
further 41 exploration wells (see Appendix). These data
are complemented by 11,500 km of seismic lines and a
general coverage by aeromagnetic surveys of the south-
western, western, and northern Chaco.
There are four subbasins in the Paraguayan Chaco
(Figure 3): (1) the Curupaity subbasin in the north,
containing 3 hydrocarbon test wells; (2) the Carandaity
subbasin in the west, with 28 hydrocarbon test wells; (3)
the Pirity subbasin in the southwest, with 8 hydrocarbon
wells; and (4) the Pilar subbasin in the south, which has
no test wells to date. These four subbasins were periodi- Figure 1—Map of South America showing the location of
cally yoked together to form the broader Chaco basin the Paraguayan Chaco basin and its regional geologic
complex. setting.
The Curupaity and Carandaity subbasins are essen-
tially Paleozoic depocenters, while the Pirity and Pilar
subbasins are attributed to Mesozoic extension. The The Phanerozoic Chaco basin is characterized strati-
Paleozoic–Mesozoic San Pedro low to the east is a graphically by three episodes and styles of sedimentation
westward extension of the Paraná basin and has no wells (Figures 4, 5). First, clastic and carbonate sedimentation
on the Chaco side. Finally, the Paleozoic Bahia Negra from the latest Proterozoic to Early Permian occurred in
platform to the northeast, also without wells, is largely marine and continental environments on a platform and
interpretative (Figure 3). These depocenters are sep- locally subsiding basins. Second, terrigenous clastic and
arated by intervening arches. carbonate sedimentation of Late Jurassic–Early Creta-
The Chaco basin (Figure 4) is a modern foreland basin ceous to middle Eocene age formed thick continental fills
between the Andean ranges to the west and the Brazilian in rift basins, with local marine transgression. Third,
shield to the northeast. Toward the east and south, it terrigenous clastic and evaporitic sedimentation with
merges with the Paraná and Pampa basins, respectively. local marine incursions occurred from the middle Eocene
The tectonic style of the Chaco basin is characterized by to Quaternary throughout the Chaco basin in a foreland
northwest- and northeast-oriented structural lineaments basin setting.
of Eocambrian Brasiliano cycle origin. Differential reacti-
vation of these fabrics through Phanerozoic time resulted
in four distinct phases of subsidence: early Paleozoic, late GEOLOGIC SETTING
Paleozoic, late Mesozoic, and Cenozoic (Figure 5). The
phases are separated by erosional unconformities or The tectonic history of the Phanerozoic Chaco basin
marked by nondeposition or low sedimentation rates. started during the intense thermotectonic Brasiliano cycle
While the Paleozoic phases reflect mild subsidence and (680–450 Ma), when carbonate and clastic sequences of
local structural readjustments, the Mesozoic basins were the Eocambrian Itapucumí Group were deposited. The
subjected to a general reorganization of the structural Brasiliano basins evolved through extensional and
styles by extension along predominantly northeast- compressional phases (Zalan, 1987). Thrusting along the
oriented lineaments related to the Atlantic opening. edges of the basins and acid magmatism (680–580 Ma)
Finally, a Cenozoic phase was caused by the Andean affected southeastern and northeastern Paraguay
orogeny and accompanying regional structural adjust- (Cordani, 1984; Wiens, 1986). This Brasiliano event marks
ments. the initiation of the Chaco and Paraná basin subsidence
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 187

Figure 2—Simplified geology of Paraguay. The study area of the Paraguayan Chaco basin is on the western side of the
Paraguay River.

(Zalán, 1987) and establishment of the northwest- Paleozoic


southeast and northeast-southwest structural framework
(Figure 6). For much of the Paleozoic, the Chaco basin was part
This complex tectonic fabric is expressed as basement of a relatively stable area of shallow marine and conti-
highs such as the Río Apa and Río Tebicuary subcratonic nental sedimentation south of the Brazilian shield
blocks, platforms on which the carbonate Itapucumí (Figures 1, 5). The Paleozoic succession is dominated by
Group was deposited, and local areas of subsidence in terrigenous clastics.
the Carandaity and Curupaity regions (Figure 5). An almost complete Ordovician sequence of the Cerro
188 Wiens

Figure 3—Major Cretaceous tectonic units of Paraguay


showing the distribution of hydrocarbon exploration wells
and the locations of geologic and seismic sections
mentioned in the text. Abbreviations: Fil, Filadelfia; PV,
Puerto la Victoria; AJ, Adrian Jara; FA, Fortin Aroma; PI,
Palmar de las Islas; Me-1, Mendoza-1 well; Me-2, Mendoza-
2 well; To-1, Toro-1 well; Ga-1, Gato-1 well; Kat-1, Katerina-
1 well; Pa-1, Parapití-1 well; Dq-1, Don Quijote-1 well; Lo-1,
Lopez-1 well; Car-1, Carmen-1 well; Pal-1, Palo Santo-1
well; Be-1, Berta-1 well; Ac-1, Tte.Acosta-1 well; Naz-1,
Nazareth-1 well; Ori-1, Orihuela-1 well; Asu-1, Asunción-1
well; Asu-2, Asunción-2 well.

León Group is preserved in the Don Quixote-1 well in


the Carandaity subbasin. The Asunción-1 well in the San
Pedro low contains Middle–Upper Ordovician sedimen-
tary rocks of the Caacupé Group. Elsewhere, the succes-
sion is thin or absent, suggesting differential subsidence
and local erosion.
Outcrop geology in southeastern Paraguay and in the
northern Chaco basin suggests that continuous sedimen-
tation in Ordovician–Devonian time resulted from a
major transgressive-regressive cycle. A regional Lower
Silurian unconformity associated with the Zapla tillites in
the Chaco basin (Russo et al., 1979) has not been recog-
Figure 4—Simplified lithostratigraphic column of the
nized in the Paraguayan Chaco. Local Devonian trans- Paraguayan Chaco basin.
gression deposited the San Alfredo Group in western
and northern Chaco (Harris, 1959).
The broadscale geology suggests that a large sedimen- movement of structural basement blocks and local
tary platform was established in the Ordovician and erosion across uptilted crests where Upper Carbonif-
persisted until the Mesozoic. Only the Curupaity, erous continental glacial sediments and Lower Permian
Carandaity, and San Pedro depocenters experienced shallow marine sequences of the Palmar de las Islas
continuous subsidence, apart from the main Paraná basin. Group have an erosional contact. Locally the Permian
A pronounced angular unconformity separates the rests on Devonian shales (Lobo et al., 1976). Only in the
Devonian and Carboniferous successions in much of the deeper western parts of the Carandaity subbasin is a
northwestern Chaco. Subsequent tectonism resulted in complete section of Devonian–Upper Carboniferous and
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 189

strata are included in the lower Adrian Jara Formation. A


major Mesozoic extensional event corresponding to the
opening of the South Atlantic (230–65 Ma) is recorded for
the main Adrian Jara Formation in the Curupaity
subbasin. Its equivalents in the Pirity subbasin are the
Berta, Palo Santo, and Santa Barbara formations.
The tectonic style reflects pervasive extension from
Early Cretaceous to middle Eocene time (Figure 6).
Figures 7 and 8 show the seismic expression of these
Mesozoic rift basins. The geometry of the former
subbasins and highs was modified substantially. The
Carandaity and Curupaity subbasins became relatively
stable, with only low sedimention rates. Further uplift of
existing highs established new depositional centers.
Three new subbasins subsided along northeast-
southwest axes: the Pirity and Pilar subbasins and the
Bahía Negra platform. Continental sedimentation was
widespread. In the Pirity subbasin, a short marine
incursion of Late Cretaceous age came from the
southwest. Basic to alkaline magmatism (135–108 Ma
and about 70 Ma) is characteristic. Mesozoic rifting is
believed to have established the principal structures and
initiated maturation of potential source rocks for hydro-
carbon generation. The final configuration of the
subbasins, structural highs, and platforms was only
completed toward the end of the Atlantic phase of
extension (Zalán, 1987).
The Pirity subbasin is the best explored of the
subbasins in the Chaco. Intense en echelon faulting in
northeast-southwest and NNE-SSW directions caused
differential vertical and transverse movements, resulting
in an asymmetric half-graben structure. Local basic
magmatic activity accompanied the event.
The sedimentary infill is mostly continental, except for
periodic marine influence from the southwest as far as
the Acosta-1 well, depositing the main section of the Palo
Figure 5—West-southeast cross sections showing the Santo Formation. Sediments were deposited in alluvial
Phanerozoic evolution of the Paraguayan Chaco basin and fan, fluvial, and lacustrine settings, with local eolian
the western Paraná basin (Ordovician–Silurian to Recent tracts. The marine sediments are represented by carbon-
surface) (section A–A' on Figure 3). Note the Paleozoic ates, clastics, and evaporites. The thickness of this sedi-
migration and diversification of depocenters, as well as the mentary succession varies up to 4000 m.
Mesozoic structural reorganization and Cenozoic readjust- Toward the northeast in the Filadelfia area, the Pirity
ment and inversion features. subbasin gradually loses its identity and records minor
sedimentation. The extensional deformation shifts into
the Bahía Negra platform and the initiating Pantanal
Lower Permian sedimentary rocks preserved. In the subbasin.
Curupaity subbasin, an erosional unconformity sep-
arates Carboniferous–Permian and Mesozoic sequences. Cenozoic
Paleozoic sedimentation in the Chaco basin was
controlled by northwest and northeast Eocambrian The growth of the Andean ranges to the west of the
block-type structural trends. Vertical and horizontal Chaco basin (50–35 Ma) created a new source for
movements were generally minor, but nevertheless suffi- sediments and essentially excluded marine influence
cient to form depocenters and intrabasinal highs, as well from that direction. Tectonic and magmatic responses
as controlling the distribution of sedimentary facies. are recorded in the Asunción and San Ignacio blocks
along the edge of the basement in eastern Paraguay as
Mesozoic inverted rift basin remnants with local nephelinitic
magmatism (49–40 Ma).
Fluvial and eolian sedimentation during the Triassic The subbasins and highs of the Chaco basin that were
filled depressions (Bianucci et al., 1981). In the northern established during the Mesozoic became covered by the
Chaco, the environment changed gradually to a desert thick continental cover of the Chaco Formation. Slight
landscape, blanketing the remaining topography. These reactivation of structures in the Mesozoic are attributed
190 Wiens

Figure 6—Southern Paraguay


showing regional tectonosedi-
mentary framework (thickness
contours in meters). Abbrevia-
tions: NpE, Río Apa subcraton;
SpE, Río Tebicuary subcraton;
Be-1, Berta-1 well; An-1, Anita-1
well; Car-1, Carmen-1 well; Pir-1,
Pirizal-D1 well; Gl-1, Gloria well;
Pal-1, Palo Santo-1 well; Ac-1,
Acosta-1 well; Naz-1, Nazareth-1
well; Asu-1, Asunción-1 well;
Asu-2, Asunción-2 well; FMB-1,
Mariano Boedo-1 well (Argentina);
Pr-1, Pirané-1 well (Argentina);
PL-1, Palmar Largo-1 well
(Argentina).

to Tertiary tectonism. That some of these structures are The Itapucumí Group is subdivided into a basal trans-
still active is evidenced by internal drainage patterns in gressive unit (shale, arkosic sandstone, and conglomerate
the Bahía Negra–Pantanal area and weak seismic activity. beds up to 25 m thick) and a major calcareous sequence
(Wiens, 1986). The carbonate interval consists of bitumi-
nous and laminated limestones that alternate with
PHANEROZOIC STRATIGRAPHY abundant oolitic and conglomeratic beds and
interbedded shales, parts of which show grades of meta-
The Phanerozoic succession in the Paraguayan Chaco morphism. The Itapucumí Group is 250–400 m thick.
ranges from Eocambrian to Recent. This interpretation is Biostratigraphic dating is based on algal remains and
based on outcrops in the northern Chaco basin and on Scyphozoae remnants and suggests a latest Proterozoic–
exploration data and is accompanied by outcrop and Cambrian age (Beurlen and Sommer, 1957; Correa et al.,
well data from eastern Bolivia (e.g., Tucavaca, Roboré, 1979; Hahn et al., 1982; Hahn and Pflug, 1985; Aceñolaza
and Santiago de Chiquitos), outcrop and well data from et al., 1989). These deposits were generated by transgres-
the sub-Andean belt of Bolivia and Argentina, hydro- sion over a continental platform, resulting in a shallow,
carbon exploration data from the Argentinian Chaco, warm marine environment.
and geologic data from eastern Paraguay (Figures 2, 9). The Cerro León Group is of Early Ordovician–Silurian
The Precambrian basement beneath the Chaco is poorly age (Figures 4, 9). It crops out in the Cerro León massif in
understood. Nevertheless, it is believed to resemble the the northern Chaco basin (Wiens, 1991) and in the
surrounding Río Apa and Río Tebicuary subcratonic Cordillera de los Altos area of eastern Paraguay. Only
blocks of eastern Paraguay or the Brazilian shield of one exploration well in the Carandaity subbasin has
eastern Bolivia. reached Lower Ordovician strata (Parapití-1 well)
(Figures 11a, b). This section represents a transgressive-
Paleozoic regressive cycle linked to the geodynamic evolution of
the early Paleozoic Pacific margin.
The Itapucumí Group is of latest Proterozoic–Cambrian Although the lowest Ordovician levels reached in the
age. Outcrops occur along the Paraguay River in the area Chaco basin contains dark marine shales and siltstones
of Puerto La Victoria. It is correlative to Eocambrian sedi- with lingula (Vistalli, 1989), a conglomeratic unit with
mentary rocks of the Corumbá Group in Mato Grosso do skolithos is its equivalent in the Cordillera de los Altos.
Sul, Brazil, and the Tucavaca Group in eastern Bolivia The lowest Ordovician is the La Paz Formation (Wiens,
(Wiens, 1986) (Figure 10). 1989), which in the northwestern Chaco is a black, pyritic
Figure 7—Segment of a north-south seismic line showing the northwestern flank of the Pirity subbasin and subsidence along listric
normal faults (southwestern Paraguayan Chaco basin; section D–D' on Figure 3).
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential

Figure 8—Segment of a north-south seismic line showing the southeastern flank of the Pirity subbasin and disposition along a
normal discordance (southwestern Paraguayan Chaco basin; section E–E’ on Figure 3).
191
192 Wiens

Figure 9—Comparative stratigraphic chart of the Paraguayan Chaco basin, western Paraguay, with adjacent correlative
areas in eastern Paraguay, Argentina, and Brazil.

Figure 10—Stratigraphy of the Eocambrian along the Río Apa subcraton and the southern border of the Guaporé shield in
Paraguay, Brazil, and Bolivia.

shale interval with interbedded siltstones and sand- and 335 m (Parapiti-1 well) (Clebsch, 1991). An open-file
stones. Don Quixote-1 well penetrated 172 m of this Pennzoil–Victory Oil (1972) report records crinoid stems
shaley sequence (Clebsch, 1991). A low-energy lagoon in the shales. Wolfart (1961) reports arthrophycus,
environment is inferred. The entire sequence is an brachiopods, and gastropods from the coarser intervals
upward-coarsening progradational unit deposited in a near the top. A Llandoverian age is suggested.
restricted environment. The Ordovician–Silurian Cerro León Group is wide-
Marine regression persisted throughout the Silurian, spread, blanketing the Paraguayan Chaco eastward
depositing thick sandstones with shaley interbeds. These (López-1 and Orihuela-1 wells), across the San Pedro low
gradational deposits are the uppermost La Paz formation (Asunción-1 and Asunción-2 wells), and reaching the
and the Santa Rosa Formation (Wiens, 1989). Measured main Paraná basin (Wood and Miller, 1991). Isopach
thicknesses vary between 135 m (Don Quixote-1 well) maps of the lower Paleozoic sequence (Figures 11a, b)
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 193

Figure 11—Depth and isopach maps (contours in meters).


(a) Ordovician–Silurian depth contour map of the
Paraguayan Chaco basin. (b) Ordovician–Silurian isopach
map of the Paraná basin. (c) Devonian isopach map of the
Paraguayan Chaco basin and the Paraná basin.
(d) Carboniferous isopach map of the Paraguayan Chaco
and Paraná basins. (e) Mesozoic sediment isopach map of
the Paraguayan Chaco and Paraná basins.
194 Wiens

Figure 13—Stratigraphy of the Curupaity subbasin,


northern Paraguayan Chaco basin. (Adapted from Lopez
Figure 12—Simplified stratigraphy of the Carandaity Paulsen et al., 1982; Wiens, 1989.)
subbasin, western Paraguayan Chaco basin. (Modified
from Clebsch, 1991.)
et al., 1982). Continental sedimentation was restricted to
the central Chaco uplift (López-1 and Orihuela-1 wells).
and stratigraphic correlations confirm a regional marine The lower San Alfredo Group reflects shore zone to
influence throughout the Chaco and Paraná basins. The shallow marine environments (Fernandez Garrasino and
various depocenters were yoked together forming a Cerdan, 1981). White, cross-bedded quartz arenites and
broad epeiric basin. less mature sandstones up to 300 m thick contain interca-
The San Alfredo Group is of Early–Late Devonian age lated siltstones and shales, indicating a variety of sedi-
(Figure 9). Extensive Devonian outcrops occur in the mentary facies. The presence of leiospheres and chitino-
Cordillera de San Alfredo and along the western margin zoans in the northwestern Chaco exploration wells
of the Cerro León massif. Almost all exploration wells in (Pennzoil–Victory Oil, 1972) and corals, bryozoans, and
the western and northern Chaco within the Carandaity crinoids at the Cerro León massif are indicative of a
and Curupaity depocenters have encountered Devonian marine depositional environment. Wolfart (1961) inter-
sections, as have groundwater surveys throughout the preted Favosites sp., Leptocoelia flabellites, Chonetes falk-
northern Chaco (PNUD, 1991) (Figure 11c). landicus, and Tentaculites stubeli at Fortin Aroma and the
The thickest Devonian sections occur in the Lagerenza high as evidence for an Early Devonian age.
Carandaity (3500 m) (Figure 12) and Curupaity (2000 m) Terrestrial spores at the central Chaco uplift show that
(Figure 13) depocenters. They include a lower sandy sandy coastal facies rimmed the arched areas; basinward
interval and a monotonous shaley unit at the top that the facies are deeper marine (Wiens, 1991). Postdeposi-
grades upward into a more arenaceous interval. tional diagenesis, lateral sedimentary variations, and
A generally transgressive sea inundated the Chaco- intraformational unconformities characterize reservoir
Paraná basin complex in Devonian time (Lopez Paulsen, intervals.
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 195

The upper San Alfredo Group is represented by dark, Mesozoic–Cenozoic


thick, fossiliferous, shallow marine shales (Lopez
Pugliessi and Suarez-Soruco, 1982). It grades upward The Adrian Jara Formation of Late Jurassic–middle
into a regressive sequence of continental deposits. Late Eocene age (Figures 11e, 14) is a poorly sorted sandstone
San Alfredo shale deposition were restricted largely to with interbedded conglomerates and mudstones. It
the Curupaity and Carandaity depocenters. Tentaculites, occurs principally in the Curupaity subbasin, although
brachiopods, and crinoids indicate shallow marine facies there are broadly contemporaneous deposits in the
(Harrington, 1956). Subsequent tectonism and burial northern and northwestern Chaco.
maturation have resulted in Devonian shales that range The Adrian Jara is separated from the Carboniferous
from immature (Carandaity and Curupaity subbasins) to Cabrera Formation below and from the Cenozoic Chaco
low-grade metamorphic (Lagerenza and Fuerte Olimpo Formation above by unconformities. Although this
highs). formation is imprecisely dated, there is evidence that
The uppermost regressive sedimentary rocks range deposition started in the Late Jurassic, or possibly even
in thickness up to 300 m. The paleontology records a earlier (Toro-1 well). The Adrian Jara consists of
pronounced hiatus between these deposits and the medium- to fine-grained sandstones that are moderately
overlying Carboniferous sequence (Lopez Paulsen et al., sorted, horizontal, and cross bedded. Heterogeneous
1982). Depositional environments were similar to those conglomerates overlie local scour surfaces. Each deposi-
of the lower San Alfredo Group, although lateral facies tional sequence is interbedded with thin mudstones.
changes show a strong continental influence. The Fluvial and eolian depositional environments are inter-
deposits record the termination of the Devonian trans- preted.
gressive-regressive cycle in the Paraguayan Chaco The Upper Jurassic–middle Eocene stratigraphy of the
basin. Pirity subbasin comprises three formations: the Berta,
The Palmar de las Islas Group is of Early Carbonif- Palo Santo, and Santa Barbara (Figure 14). Mesozoic
erous–Early Permian age. Significant Carboniferous extensional tectonism formed the Pirity and Pilar
outcrops occur in the northern Chaco along the subbasins and structurally modified the Bahía Negra
western Lagerenza high and in the area from Palmar platform (Figures 15, 16). The asymmetric half-graben
de las Islas to Adrian Jara. Favorable reservoir and has pronounced step faults on the northern flank
aquifer characteristics have resulted in numerous wells (Boquerón high) and a structural flexure along the
being drilled throughout the northwestern Chaco southern flank (Presidente Hayes uplift). Interpretation is
(Carandaity and Curupaity subbasins). The Carbonif- based on exploration wells. Lithologic correlations, pale-
erous isopach map reconstruction is shown in Figure ontologic data, and radiometric ages from intrusives
11d. The succession is up to 1600 m thick and is suggest local subdivisions (Salfity and Marquillas, 1981)
separated from the Devonian San Alfredo Group by a (Figure 14).
prominent unconformity (Sanjines, 1982). Depositional The Upper Jurassic–Upper Cretaceous Berta
environments included shallow marine, continental, Formation in the Pirity subbasin represents the early
and glacial conditions. stage of rift subsidence (Yacimientos Petroliferos Fiscales,
The Lower Carboniferous unit is the San José 1984) (Figure 14). The sediments where deposited in
Formation, which is characterized by rapid lateral eolian, alluvial fan, braided river, and lagoonal environ-
changes in sedimentary facies with only local continuity. ments. Equivalents of the Berta Formation are shown in
Basal sandstones with dark shaley lenses are replaced Figure 9. The Berta Formation consists of fine-grained
upward by varved mudstones and reddish diamictites reddish sandstones with interbedded mudstones. Its
(Lopez Paulsen et al., 1982). While coastal marine condi- thickness varies up to about 3000 m. Correlation with
tions are interpreted for the basal section, continental Argentinian equivalents suggests a Coniacian age (Carle
glacial environments dominated the upper levels. Local et al., 1991). Absolute ages from basaltoids indicate an
discordances and variable thicknesses (up to 800 m in the Early Cretaceous (Valanginian) age for the formation
Carandaity subbasin) are characteristic. (128 ± 5 to 126 ± 3.5 Ma) (Galliski and Viramonte, 1985).
The Upper Carboniferous–Lower Permian Cabrera The basalts were emplaced along northwest-southeast
Formation has a transitional contact with the underlying and northeast-southwest fault trends.
San José Formation and is a sandstone-dominated The overlying Palo Santo Formation is of Late Creta-
sequence with local basal conglomerates. The sandstones ceous–early Paleocene age. It records transgressive
are typically cross-bedded units fining upward into flooding of the Pirity subbasin from the southwest
mudstones with oolitic limestone levels. Shallow marine (Moreno, 1970). The Palo Santo Formation is commonly
to fluvial depositional processes are inferred (Lopez subdivided into three units (Figure 14):
Paulsen et al., 1982). Reasonable porosity and perme-
ability parameters are laterally variable. The Cabrera 1. The transgressive lower Palo Santo Formation
Formation in the Curupaity subbasin ranges in thickness sandstones onlap the margins of the Pirity
up to 1300 m. subbasin. The sandstones were deposited in
Biostratigraphic relationships of the Palmar de las alluvial fan and braided river plain environments
Islas Group are not firmly established because only rare (Turner, 1959). A Campanian–Maastrichtian age
fossils are reported from lower Upper Carboniferous red for the transgression is indicated by seismic and
beds (Lobo et al., 1976). lithostratigraphic correlations with northern
196 Wiens

Argentinian fossil-dated sequences (Reyes and


Salfity, 1973; Bonaparte et al., 1977) and by absolute
ages from basic extrusives at the base of the lower
Palo Santo Formation (70 ± 5 Ma, K/Ar) (Carle et
al., 1989). Maximum thickness is 200–220 m.
2. The middle Palo Santo Formation has a gradational
base with the lower transgressive interval. It
records maximum flooding of the Late Creta-
ceous–early Paleocene sea from the southwest
(Figure 15). At the base, littoral clastics and carbon-
ates interfinger with fluvial and eolian sandstones.
With progressive flooding, a carbonate-shale
platform with local evaporites developed (Gómez
Omil et al., 1989). Comparison of well sections
across the Pirity subbasin between the Berta-1 and
Palo Santo-1 wells (north-south section), and
between the Carmen-1 and Nazareth-1 wells
(southwest-northeast section), indicates argilla-
ceous calcite-dolomite cemented sequences in the
deeper parts and coarser clastics along the margins
of the subbasin. Sporadic ostracods and paly-
nomorphs in Berta-1 well support an interpretation
of shallow marine and coastal plain environments
in the early Paleocene. Farther away in northern
Argentina, Late Cretaceous foraminifera and
ostracods suggest shallow marine and even fresh-
water environments (Mendez and Viviers, 1973;
Moroni, 1982).
In summary, in middle Palo Santo time, the Pirity
subbasin was a restricted carbonate basin with
fringing brackish lagoonal, fluvial, and eolian envi-
ronments (e.g., Acosta-1 and Nazareth-1 wells). It is
a finely laminated succession indicating overall
low-energy conditions. Thickness varies between
200 and 550 m.
3. The upper Palo Santo Formation is a regressive
unit containing 100–230 m of evaporites,
interbedded mudstones, and marginal sandstones.
Halite, gypsum, and dolomite banks indicate
typical lagoonal environments with fringing sandy
fluvial margins (Gómez Omil et al., 1989).

The lower Paleocene–middle Eocene Santa Barbara


Formation consists of red mudstones, evaporites, and
local carbonates that are attributed to fluvial and lacus-
trine environments in a desiccated Pirity subbasin. Like
the previous unit, it has a threefold subdivision (Figure
14) (Pascual, 1978):

1. The lower Santa Barbara Formation is up to 800 m


thick and consists of reddish mudstones with
subordinate sandstones and evaporites. Paly-
nomorphs in Palo Santo-1 well indicate a Paleocene
age (Gómez Omil et al., 1989).
2. The middle Santa Barbara Formation is up to 190 m
thick and consists of claystones, marls, siltstones,
and sporadic sandstones. Lacustrine environments
Figure 14—Lithostratigraphic chart of the are inferred (Millioud, 1975).
Cretaceous–Tertiary in the Pirity, Carandaity, and 3. The upper Santa Barbara Formation is up to 1100 m
Curupaity subbasins of southwestern, western, and thick, representing an early–middle Eocene cover
northern Paraguayan Chaco basin. of the Pirity subbasin (Pascual, 1978; Quattrocchio,
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 197

Figure 15—Seismic time contour distribution of the Upper Cretaceous–Lower Paleocene Palo Santo Formation in the Pirity
subbasin, southwestern Paraguayan Chaco basin (contours in seconds; l, hydrocarbon exploration wells). (Adapted from
Clebsch, 1991.)

Figure 16—Structural and stratigraphic profile of southwest-southeast Paraguay (section B–B’ on Figure 3).

1980). The succession drapes the structural margins stones with thinly interbedded siltstones and fine-
of the Pirity subbasin. A period of regional or grained sandstones. A regional “green horizon”
epeiric basin type of subsidence is indicated. Lacus- marker bed that originated in a lacustrine setting is
trine and local ephemeral fluvial environments characteristic (Gómez Omil et al., 1989). It is a 2–10-
persisted. These depositional environments have m-thick calcareous shale and oolitic limestone with
resulted in a monotonous succession of red clay- gypsum levels.
198 Wiens

Godoy, 1990). Stratigraphic relationships suggest an


internal drainage basin dominated by ephemeral braided
fluvial systems, alluvial fans, playa lakes, and eolian
dunes. The cross-sectional geometry (thickness of 800 m
in the west, 250 m in the east) is typical of a foreland
basin.
The Quaternary (early Pleistocene–Recent) deposits of
the Chaco basin reflect a continuation of the Chaco
Formation system with heterogeneous depositional para-
meters (Bertoni, 1939; Hoffstetters, 1978; Presser and
Crossa, 1984; Herbst and Santa Cruz, 1985).

HYDROCARBON EXPLORATION
POTENTIAL
Hydrocarbon potential in the Paraguayan Chaco is
related to the Paleozoic and Mesozoic marine shales and
carbonates (Figure 17). Although no commercial hydro-
carbon discoveries have yet been made, the Chaco
remains a relatively unexplored frontier region.
Hydrocarbon seeps occur in the Tucavaca area of
eastern Bolivia in lowermost Phanerozoic Itapucumí
Group limestones and shales. Surface gas indications are
reported from correlative bituminous sedimentary
deposits in northeastern Paraguay and Goias (Brazil).
Paleozoic marine shales and carbonates reach thick-
nesses of 2500 and 3600 m in the Carandaity and
Curupaity subbasins, respectively. Organic material in
the upper parts of the succession in the least deformed
interior parts of the subbasins (e.g., Katarina-1 well)
tends to be immature. In the deeper parts of the section
and along the margins of the subbasins, the level of
organic maturity ranges from mature to overmature
(e.g., Don Quixote-1 well). The increasing geothermal
Figure 17—Summary of the Paraguayan stratigraphy and maturity with depth results in decreasing gas wetness.
its hydrocarbon potential. l, source rock; m, reservoir
Similar conditions occur along structural highs where
rock.
increasing geothermal gradients (Figure 18), fracturing,
and contemporaneous magmatism have matured the
The middle Eocene–early Pleistocene Chaco Formation sedimentary rocks, even to low-grade metamorphic
is a 500–1000-m-thick cover succession demonstrating conditions (e.g., Toro-1 well).
that the various depocenters of the Paraguayan Chaco The part of the sequence with the most hydrocarbon
had lost their individual identities. They were replaced in potential is undoubtedly the Devonian San Alfredo
the early Tertiary by an overfilled regional downwarp. shales. The Lower Ordovician Cerro León shaley section
The Andean orogen became the major supplier of also has some potential; in Argentina and Bolivia there
sediments. are producing fields in this interval. Dry gas (Mendoza-1
The lower Chaco Formation of middle Eocene–late and Mendoza-2 wells) and oil shows (Toro-1 well) are
Pliocene age was dominated by continental deposition noted. High-gravity oil most likely occurs in the upper
(Mingramm et al., 1979), which was interrupted in the parts of the sections in the interior of the subbasins, but it
southeastern Chaco by a short interlude of shallow probably changes to wet gas, condensate, and even dry
marine transgression (middle Miocene) (Russo and gas in the lower sections. Structural highs may be gas
Chebli, 1979). This marine incursion from the Atlantic prone.
(Chebli et al., 1989) reached as far as the Boquerón high Mesozoic transgressive marine deposits of the Palo
and the central Chaco uplift. Santo Formation are a local hydrocarbon play in the
The late Pliocene–early Pleistocene upper Chaco Pirity subbasin. Gas and oil indicators in the Berta and
Formation comprises alternating fine-grained sandstone, Palo Santo formations suggest potential for smaller oil
siltstone, and claystone. The succession coarsens to the concentrations (Schlumberger, 1987; Fernandez Garrasino,
west, toward the growing Andean orogen. These 1989). This interval contains producing fields in northern
sediments represent the most important aquifers in the Argentina. The sedimentary distribution, local
Paraguayan Chaco (Tullstrom, 1973; Osterbaan, 1988; magmatism, and geothermal gradients are restrictive.
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 199

fractured, intruded by magmatites, and have geothermal


gradients up to 0.6 C˚/100 m. The most favorable condi-
tions probably exist along the flanks of the highs. The
interior of the Curupaity subbasin is expected to be
largely immature.
The best quality reservoir rocks are the Lower
Carboniferous San José Formation and Upper Carbonif-
erous Cabrera Formation channel sandstones, which
resemble some of the producing reservoirs of the
adjacent Bolivian Chaco basin. Excellent reservoir charac-
teristics also occur in the Upper Jurassic–middle Eocene
Adrian Jara Formation. The Mesozoic sandstones may
have poor seals. Structural traps include Tertiary
inversion of Mesozoic extensional fabrics.
The Devonian upper San Alfredo Group shales in the
Carandaity subbasin (Figure 5) exceed 2500 m in thickness.
They represent excellent hydrocarbon source rocks. Early
Ordovician Cerro León Group shaly units (TOC content
of 0.5 wt. %; Don Quixote-1 well) may represent a
secondary source rock. Exploration targets include inter-
calated uppermost Devonian sandstones and wide-
spread channel sandstones of the Carboniferous Palmar
de las Islas Group.
The flanks of basin margin structural highs have
mature organic facies. Very mature conditions occur on
the crests of the highs (e.g., Mendoza-1 and Mendoza-2
wells) where there are gas shows. The depocenters are
immature (e.g., Katerina-1 well). Eocene shale-draped
Figure 18—Geothermal gradient pattern in the north-
unconformities act as seals along the Lagerenza and
western Paraguayan Chaco basin (contours in ˚C/100 m). Boquerón highs, as well as toward the Central Chaco
uplift.
The principal hydrocarbon potential in the Pirity
The regional distribution of geothermal gradients in subbasin (Figure 14) is the Upper Cretaceous marine
the Paraguayan Chaco (calculated on surface tempera- source rock of the Palo Santo Formation. Geochemical
tures of 37.7˚C) traces contours that follow the main assays indicate oil-generative conditions in the shales
structural units (Figure 19), indicating that sedimentary and carbonates of the middle Palo Santo Formation
maturity was mostly influenced by tectonism. Geo- (Carle et al., 1991). Reservoir intervals include the sand-
thermal gradients over structural highs may be as high stones of the middle Palo Santo or the underlying lower
as 0.7–0.9˚C/100 m (Lagerenza and Fuerte Olimpo Palo Santo and Berta formations. Brecciated volcanics
highs), possibly reflecting local magmatic activity as intercalated in the Berta and Palo Santo formations may
indicated by magnetics and outcrop studies (Cerro León offer viable prospects, just as the Palmar Lago field in
massif). The Toro-1 well has fractures and volcanics as northern Argentina does.
compared to the Gato-1 well, which has no volcanics and The western part of the San Pedro low (Figure 19)
lower thermal gradients. High levels of organic maturity reaches into the Paraguayan Chaco from the Paraná
and tight reservoir sandstones characterize the structural basin with a continuation of Phanerozoic deposits. It
highs. contains Lower Ordovician, Silurian, Devonian, and
Overall, the various subbasins have low geothermal Permian source rocks.
gradients with maturation affecting the lower sedimen- On the Bahia Negra platform (Figure 5), oil seeps, gas
tary sequences. Basinal structural deformation is not shows, and high bituminous kerogen content indicate a
significant. On the basis of sedimentary and diagenetic hydrocarbon source potential in lower Phanerozoic
history, tectonism, magmatism, and geothermal Itapucumí Group sedimentary rocks of the northeastern
gradients, it is concluded that the Paleozoic section of the Chaco. Although there is structural trap potential,
western and northern Chaco has the highest hydro- reservoir quality is expected to be poor. Nevertheless,
carbon potential; Mesozoic plays are concentrated in the this remains an unexplored area.
southwestern Chaco. In the northern part of the Pilar subbasin, marine
In the Curupaity subbasin (Figure 3), the 3600-m-thick clastics of Ordovician–Silurian age have built a Paleozoic
upper San Alfredo shales represent an important hydro- terrace wedge along the margin of the Río Tebicuary
carbon target; an oil show in the Toro-1 well has TOC subcraton in the southeastern Chaco. These sedimentary
values between 0.3 and 2.1 wt. %. Slatey shales in the rocks are covered by Cretaceous–Paleocene synrift half-
Toro-1 well indicate overmature conditions along the graben fills of continental origin. Exploration efforts in
Fuerte Olimpo high, where the rocks are highly Argentina have so far been negative for hydrocarbons.
200 Wiens

Figure 19—North-south structural and stratigraphic profile of eastern Paraguay (section C–C ' on Figure 3).

CONCLUSIONS marine and continental clastic deposits, much of


which were glacially influenced (Palmar de las Islas
The Chaco basin in Paraguay was formed on conti- Group).
nental crust where it reactivated older structural fabrics 4. Late Jurassic–middle Eocene continental clastic
of the late Precambrian–Cambrian Brasiliano event deposits filled the Carandaity and Curupaity
(680–450 Ma). This basement framework resulted in a subbasins (Adrian Jara Formation). Mesozoic
complex system of structural blocks with preferential extension in the Pirity and Pilar subbasins (Berta
fault trends oriented northwest-southeast and northeast- Formation) was accompanied by magmatism.
southwest. The tectonostratigraphic evolution of the Angular unconformities separate these Mesozoic
Phanerozoic Chaco basin reflects repeated reactivation of synrift sequences from the Paleozoic succession.
these older structural fabrics. Of paramount importance 5. A Late Cretaceous–early Paleocene marine trans-
was the relative orientation of the older structures to gression and deposition of terrigenous clastic and
prevailing stress fields. carbonate sediments are recorded in the Pirity
There were four distinct subsidence phases during the subbasin (Palo Santo Formation).
Phanerozoic: mild extension and regional epeiric basin 6. An early Paleocene–middle Eocene cover of lacus-
subsidence characterized the (1) Ordovician–Devonian trine and fluvial sediments (Santa Barbara
and (2) Carboniferous–Permian intervals. (3) Vigorous Formation) reflects a return to regional patterns of
rifting occurred in the Cretaceous–Eocene. (4) The final subsidence and burying of the remaining structural
phase from the Eocene onward was characterized by a relief.
return to mild, regionally controlled patterns of subsi- 7. A pronounced hiatus beneath a westward-thick-
dence. Each episode preserves a unique record of uncon- ening middle Eocene–early Pleistocene wedge of
formity-bounded sequences and sedimentation. continental deposits marks the onset of foreland
The Phanerozoic stratigraphic section in the basin subsidence in front of the Andean orogenic
Paraguayan Chaco ranges from latest Proterozoic to belt (Chaco Formtion). Westward-derived sedimen-
Recent in age. Crystalline basement has so far not been tation was interrupted in the southeastern Chaco by
reached by exploration drilling. The chronologic history a local marine incursion in the middle Miocene.
of events based on surface and subsurface data is as 8. Early Pleistocene–Recent sediments blanketed the
follows: Chaco basin (Quaternary).

1. Latest Proterozoic–Cambrian marine clastics and Hydrocarbon potential in the Paraguayan Chaco
limestones of the Itapucumí Group were deposited. basin is related to Paleozoic and Mesozoic marine shales
2. An erosional unconformity separates these rocks and carbonates in each of the sedimentary subbasins and
from Early Ordovician–Silurian and Early to the way these sedimentary sequences abut the struc-
Devonian marine shales and sandstones of the tural highs. The Lower Ordovician Cerro León Group
Cerro León and San Alfredo groups. and Devonian upper San Alfredo Group shales are
3. An angular unconformity intervenes between these source rocks in the Carandaity and Curupaity subbasins.
and Early Carboniferous–Early Permian shallow Oil generation is inferred for the lower parts of the
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential 201

sequence in the center of each subbasin and along the Chebli, G. A., O. Tofalo, and G. E. Turzzini, 1989,
flanks of the basin margin highs. Prospective reservoir Mesopotamia, in G. Chebli and L. Spalletti, eds., Cuencas
intervals include Devonian intraformational sandstones, sedimentarias argentinas: Serie Correlación Geológica,
Carboniferous channel sandstones, and Mesozoic sand- Universidad Nacional de Tucumán, no. 6, p. 79-100.
Clebsch, C. A., 1991, The geological evolution of the
stones. Mesozoic extension created a variety of structural
Paraguayan Chaco: Ph.D. dissertation, Texas Technical
traps. University, Austin, Texas, 185 p.
Upper Cretaceous marine transgressive deposits of Cordani, U. G., 1984, Estudo preliminar de integraçao do
the Palo Santo Formation are the most prospective units Precambriano com os eventos tectónicos das bacías sedi-
in the Pirity subbasin. Along the southeastern flank of mentares brasileiras: Boletím Ciencia-Técnica-Petróleo,
the Boquerón high, Devonian shales were block faulted Seçao: Exploraçao de Petróleo, Publicaçao, Petrobras-
during Mesozoic extension and were unconformably Cenpes-Sintep, no. 15, 28 p.
overlain by Cretaceous and Tertiary clastics and shales, Correa, J. A., F. C. L. Correia Filho, G. Scislewski, C. Neto, L.
forming viable hydrocarbon targets. A. Cavallón, N. L. S. Cerqueira, and V. L. Nogueira, 1979,
The Chaco basin of Paraguay remains largely unex- Geología das regioes centro e oeste de Mato Grosso do
Sul. Projeto Bodoquena: Ministerio das Minas e Energía,
plored. The geology suggests that it has significant
DNPM, Serie Geología Basica, v. 3, 111 p.
hydrocarbon potential. Counterparts of the Paraguayan Fernandez Garrasino, C. A., 1989, La cuenca Chaco-Paranense
Chaco basin in Bolivia and northern Argentina are major argentina: Sus tendencias evolutivas y algunas posibili-
producers of oil, gas, and condensate. dades exploratorias: Boletín de Informaciones Petroleras,
Tercera Epoca, no. 6, p. 2–17.
Fernandez Garrasino, C. A., and J. J. Cerdan, 1981, La
Formación Santa Rosa (Devónico inferior) en la Argentina
Acknowledgments I wish to acknowledge H. J. Welsink and y sus posibles equivalentes: Octavo Congreso Geológico
A. J. Tankard for suggestions to the paper, C. A. Fernandez Argentino, San Luis, actas III, p. 193–202.
Galliski, M. A., and J. G. Viramonte, 1985, The Cretaceous
Garrasino for detailed discussions, M. P. de Pella and A. J. L.
paleorift in northwestern Argentina: a petrological
de Di Sopra for typing, J. R. Britez Urdapilleta for drafting, approach: Journal of South American Earth Sciences, v. 1,
and J. D. Ray and R. Diana for technical assistance. I am espe- p. 329–342.
cially grateful to George Kronman of Amoco Production Godoy,V. E., 1990, Caracteristicas hidrogeologicas e hidro-
Company for preparing the color display of the geologic map. químicas de la región oeste del Chaco paraguayo: Publica-
ciones del Departamento de Agua para el Chaco,
Filadelfia, 147 p.
Gómez Omil, R. J., A. Boll, and R. M. Hernandez, 1989,
REFERENCES CITED Cuenca Cretácica-Terciaria del noroeste argentino (Grupo
Salta), in G. Chebli and L. Spalletti, eds., Cuencas sedimen-
Aceñolaza, F. G. , F. R. Durand, and J. Sosa Gómez, 1989, tarias argentinas: Serie Correlación Geológica, Universidad
Cuenca del Precámbrico superior-Cámbrico inferior del Nacional de Tucumán, no. 6, p. 43–64.
noroeste argentino, in G. Chebli and L. Spalletti, eds., Hahn, G., and H. D. Pflug, 1985, Die Llondinae nova familiae,
Cuencas sedimentarias argentinas: Serie Correlación Kalk-Röhren aus dem Vendium und Unterkambrium:
Geológica, Universidad Nacional de Tucumán, no. 6, Senckenbergiana lethaea, v. 65, p. 413–431.
p. 9–18. Hahn, G., R. Hahn, O. H. Leonardos, H. D. Pflug, and D. H. G.
Bertoni, A. de W., 1939, Informe sobre rocas conchilianas de Walde, 1982, Körperlich erhaltene Scyphozoen Reste aus
Villeta: Revista de la Sociedad Científica del Paraguay, dem Jungpräkambrium Brasiliens: Geologica et Paleonto-
v. 4, 1 p. logica, v. 16, p. 1–18.
Beurlen, K., and F. W. Sommer, 1957, Observacoes estratigrá- Harrington, H. J., 1956, Paraguay, in W. F. Jenks, ed.,
ficas e paleontológicas sobre o calcareao Corumbá: Boletím Handbook of South American geology: GSA Memoir 65,
Divisáo de Geologia e Mineralogía, no. 168, p. 1–35. p. 99–114.
Bianucci, H. A., O. M. Acevedo, and J. J. Cerdán, 1981, Harris, J. W., 1959, Sample and core descriptions of López-1
Evolución tectosedimentaria del Grupo Salta en la well, in The Gran Chaco concession of Pennzoil de
subcuenca Lomas de Olmedo (provincias de Salta y Paraguay S. A.: Pennzoil unpublished report.
Formosa): Octavo Congreso Geológico Argentino, San Herbst, R., and J. N. Santa Cruz, 1985, Mapa litoestratigráfico
Luis, actas III, p. 159–172. de la provicia de Corrientes: D’orbigniana, v. 2, 50 p.
Bonaparte, J. F., J. A. Salfity, G. E. Bossi, and J. Powell, 1977, Hoffstetter, R., 1978, Une faune de mammiferes pleistocenes
Hallazgo de dinosaurios y aves cretácicas en la Formación au Paraguay: Societé geologiqué du France, fascsimile 1,
de El Brete (Salta), próximo al límite con Tucumán: Acta p. 32–33.
Geológica Lilloana, v. 14, p. 5–17. Lobo, J., M. Suarez Riglos, and R. Suarez-Soruco, 1976,
Carle, R. J.,O. E. Di Persia, and G. A. Olivieri, 1989, Análisis Nuevas unidades cronoestratigráficas para el Paleozoico
geológico y petrolero del sector noroeste de la provincia de medio de la Provicia Austral Sudamericana: Duodecima
Formosa, República Argentina: Primer Congreso Nacional Reunión de Expertos Arpel, Buenos Aires, comunica-
de Exploración para Hidrocarburos, Mar del Plata, p. ciones.
163–185. Lopez Paulsen, O., G. Montemarro-Díaz, and H. Trujillo-
Carle, R. J.,O. E. Di Persia, and H. Belotti, 1991, Geología del Ikeda, 1982, Estratigrafía del Paleozoico medio en las
sector noroeste de la provincia de Formosa: Boletín de serranías de San José y Santiago de Chiquitos, Bolivia:
Informaciones Petroleras, Tercera Epoca, v. 8, no. 26, Quinto Congreso Latinoamericano de Geología, Buenos
p. 2–17. Aires, actas I, p. 283–292.
202 Wiens

Lopez Pugliessi, J. M., and R. Suarez-Soruco, 1982, Síntesis Salfity, J. A., and R. A. Marquillas, 1981, Las unidades estrati-
estratigráfica del Devónico boliviano en la cuenca gráficas cretácicas del norte de la Argentina, in W.
subandina del sur: Quinto Congreso Latinoamericano de Volkheimer and E. A. Musacchio, eds., Cuencas Sedimen-
Geología, Buenos Aires, actas I, p. 267–282. tarias del Jurasico y Cretacico de America del Sur, Buenos
Mendez, I., and G. Viviers, 1973, Estudio micropaleontológico Aires, v. 1, p. 303-317.
de sedimentitas de la Formación Yacoraite (provincias de Sanjines, S. G., 1982, Estratigrafía del Carbónifero, Triásico y
Salta y Jujuy): Quinto Congreso Geológico Argentino, Cretácico boliviano en el borde oriental de las sierras
Buenos Aires, actas III, p. 467–470. subandinas centrales: Quinto Congreso Latinoamericano
Millioud, M. E., 1975, Palynological study of cuttings and de Geología, Buenos Aires, actas I, p. 236–264.
side-wall core samples from the Palo Santo-1 well: Conces- Schlumberger, 1987, Evaluación de formaciones en la
sion Exxon Paraguay, unpublished report. Argentina: Schlumberger/YPF, publicaciones, Buenos
Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979, Aires, 44 p.
Sierras Subandinas, in J. C. M. Turner, ed., Geología Tullstrom, H., 1973, Investigation of groundwater resources in
regional argentina: Academia Nacional de Ciencias, the central and northwestern Chaco: Publicaciones del
Córdoba, v. 1, p. 95–137. Departamento de Agua para el Chaco, Filadelfia, 118 p.
Moreno, J. C., 1970, Estratigrafía y paleogeografía del Turner, J. C. M., 1959, Estratigrafía del Cordón de Escaya y de
Cretácico Superior en la cuenca del noroeste argentino, con la Sierra de Rinconada (Jujuy): Revista de la Asociación
especial mención de los subgrupos Balbuena y Santa Geológica Argentina, v. 12, no. 1, p. 15–39.
Bárbara: Revista de la Asociación Geológico Argentina, v. Vistalli, M. C., 1989, Cuenca siluro-devónica del noroeste, in
25, p. 9–44. G. Chebli and L. Spalletti, eds.,Cuencas sedimentarias
Moroni, A. M., 1982, Correlación palinológica en las forma- argentinas: Serie Correlación Geológica, Universidad
ciones Olmedo y Yacoraite; cuenca del noroeste argentino: Nacional de Tucumán, no. 6, p. 19–42.
Tercer Congreso Geológico Chileno, Concepción, actas, Wiens, F., 1986, Zur lithostratigraphischen, petrographischen
p. 340–349. und strukturellen Entwicklung des Río Apa Hochlandes,
Osterbaan, A. W. A., 1988, Evaluación hidroquímica del agua Nordost-Paraguay: Clausthaler Geowissenschaftliche
subterránea en los acuíferos profundos en el Chaco oeste: Dissertationen, Heft 19, 280 p.
Publicaciones del Departamento de Agua para el Chaco, Wiens, F., 1989, Tectónica y sedimentación fanerozoica de la
Filadelfia, 27 p. Cuenca del Chaco (Paraguay): Publicaciones del Departa-
Pascual, R., 1978, Nuevos y singulares tipos ecológicos de mento de Agua para el Chaco, Boletín 1, p. 9–26.
marsupiales extinguidos de América del Sur (Eoceno Wiens, F., 1991, Geología y aguas subterráneas, Chaco-
temprano del noroeste argentino): Segundo Congreso Paraguay; región al norte de 2030’ (efectos de interacción):
Argentino de Paleontología y Bioestratigrafia, Buenos Primer Simposio de Aguas Subterraneas, actas, Asunción,
Aires, actas II, p. 27–63. p. 67–86.
Pennzoil-Victory Oil, 1972, Farmout proposal: Paraguay Wolfart, R., 1961, Stratigraphie und Fauna des älteren Paläo-
concession, in Pennzoil-Victory concession in Paraguay: zoikums (Silur, Devon) in Paraguay: Geologisches
Pennzoil unpublished report. Jahrbuch, Band 78, p. 29–102.
PNUD, 1978, Investigación y desarrollo de agua subterránea Wood, G. D., and M. A. Miller, 1991, Distinctive Silurian chiti-
en el Chaco: Publicaciones del Departamento de Agua nozoans from the Itacurubi Group (Vargas Peña shale),
para el Chaco, Filadelfia, 93 p. Chaco basin, Paraguay: Palynology, v. 15, p. 181–192.
PNUD, 1991, Perforaciones profundas y someras del norte del Yacimientos Petroliferos Fiscales, 1984, Estratigrafía del tramo
Chaco paraguayo: Publicaciones del Departamento de inferior del pozo descubridor Palmar Largo XL: Boletín de
Agua para el Chaco, Filadelfia, 55 p. Informaciones Petroleras, Tercera Época, no. 1 (2), 109 p.
Presser, J. B., and V. F. Crossa, 1984, Informe preliminar sobre Zalán, P. W., 1987, Tectónica e sedimentaçao da Bacía do
fósiles vertebrados pleistocénicos en la localidad de Paraná: Anais do Tercer Simposio Sul-Brasileiro de
Ytororó, Departamento Central: Ministerio de Educación y Geología, Curitiba, p. 38–67.
Culto, División General de Bienes Culturales, 16 p.
Quattrocchio, M., 1980, Estudio palinológico preliminar de la
Formación Lumbrera (Grupo Salta), localidad Pampa
Grande, provincia de Salta, República Argentina: Segundo
Congreso Argentino de Paleontología y Bioestratigráfica,
Buenos Aires, actas II, p. 131–149.
Reyes, F. C., and J . A. Salfity, 1973, Consideraciones sobre la
estratigrafía del Cretácico (Subgrupo Pirgua) del noroeste Author’s Mailing Address
argentino: Quinto Congreso Geológico Argentino, Fernando Wiens
Córdoba, actas III, p. 355–385.
Russo, A., R. Ferello, and G. Chebli, 1979, Llanura Chaco-
Geo Consultores
Pampeana, in J. C. M. Turner, ed., Geología regional Casilla Postal 166
argentina: Academia Nacional de Ciencias, Córdoba, v. 1, Asunción
p. 139–183. Paraguay
Appendix: CHACO BASIN (PARAGUAY): CHARACTERISTICS OF HYDROCARBON EXPLORATION WELLS

Geothermal
No. Operator Well Name and No. Location Year Subbasin or Total Depth TD: Geologic Formation Hydrocarbon BHT (°C) Gradient
Area (m) Indication (°C/100m)
1 Union Oil Co. Santa Rosa - 1 21°45' S 1947 Carandaity 2.310 m Cerro León Lower Oil show — —
61°41' W (south) Gr Ord./Sil.
2 Union Oil Co. La Paz D - 1 21°53' S 1948 Boquerón high 2.210 m Cerro León Silurian — — —
60°58' W Gr
3 Union Oil Co. Pirizal D - 1 23°03' S 1948 Pirity 3.148 m Sta. Barbara Lower Oil show — —
60°38' W Fm Tertiary
4 Union Oil Co. Picuiba B - 1 20°40' S 1949 Carandaity 2.290 m Low. San Devonian Oil show — —
61°56' W (central) Alfredo Gr
5 Union Oil Co. Orihuela B -1 23°24' S 1949 Pte. Hayes 2.046 m Cerro León Lower — — —
58°40' W uplift Gr Ord./Sil.
6 Pure Oil Co. Madrejón - 1 20°25' S 1957 Central Chaco 1.728 m Cerro León Lower Oil show 20.34 0.76
59°29' W uplift Gr Ord./Sil.
7 Pure Oil Co. Lagerenza - 1 20°00' S 1958 Lagerenza high 2.893 m Cerro León Lower Gas show 31.50 0.85
61°00' W Gr Ord./Sil.
8 Pure Oil Co. Lopez - 1 21°46' S 1959 Central Chaco 1.737 m Cerro León Lower — 16.20 0.52
59°58' W uplift Gr Ord./Sil.
9 Pure Oil Co. Mendoza - 1R 20°12' S 1959 Carandaity 3.244 m Cerro León Lower Gas blow 28.35 0.66
61°41' W (north) Gr Ord./Sil.
10 Placid Oil Co. Mendoza - 1 20°07'30" S 1966 Carandaity 794 m Up. San Upper — 12.96 0.73
61°45'20" W (north) Alfredo Gr Devonian
11 Placid Oil Co. Mendoza - 2 20°02'20" S 1967 Carandaity 1.247 m Up. San Upper Gas blow 13.32 0.50
61°52'10" W (north) Alfredo Gr Devonian
12 Placid Oil Co. Mendoza - 3 20°03'10" S 1967 Carandaity 704 m Up. San Upper — 11.79 0.67
61°53'10" W (north) Alfredo Gr Devonian
13 Pennzoil & Vict. Alicia - 1 20°57'02" S 1971 Carandaity 1.306 m Up. San Upper — 11.88 0.36
Holdings 61°48'57" W (central) Alfredo Gr Devonian
14 Pennzoil & Vict. Brigida - 1 21°18'50" S 1971 Carandaity 1.513 m Up. San Upper Oil show 12.42 0.35
Holdings 61°55'22" W (central) Alfredo Gr Devonian
15 Pennzoil & Vict. Cristina - 1 21°26'54" S 1971 Carandaity 643 m Up. San Upper — 9.09 0.30
Holdings 61°53'26" W (central) Alfredo Gr Devonian
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential

(continues)
203
204
Wiens

Appendix: CHACO BASIN (PARAGUAY): CHARACTERISTICS OF HYDROCARBON EXPLORATION WELLS (continued)

Geothermal
No. Operator Well Name and No. Location Year Subbasin or Total Depth TD: Geologic Formation Hydrocarbon BHT (°C) Gradient
Area (m) Indication (°C/100m)
16 Pennzoil & Vict. Dorotea - 1 21°17'01" S 1971 Carandaity 854 m up. San Upper — 10.08 0.34
Holdings 62°08'54" W (central) Alfredo Gr Devonian
17 Pennzoil & Vict. Emilia - 1 20°06'34" S 1971 Carandaity 1.222 m up. San Upper — 11.16 0.39
Holdings 62°07'14" W (central) Alfredo Gr Devonian
18 Pennzoil & Vict. Federica - 1 21°35'02" S 1971 Carandaity 800 m up. San Upper — 9.36 0.27
Holdings 62°11'59" W (south) Alfredo Gr Devonian
19 Pennzoil & Vict. Gabriela - 1 21°46'43" S 1971 Carandaity 1.016 m Up. San Upper Oil show — —
Holdings 62°00'02" W (south) Alfredo Gr Devonian
20 Pennzoil & Vict. Hortensia - 1 21°30'29" S 1971 Carandaity 765 m Up. San Upper Oil show 9.54 0.31
Holdings 61°39'27" W (south) Alfredo Gr Devonian
21 Pennzoil & Vict. Isabel - 1 21°01'14" S 1971 Carandaity 946 m Up. San Upper — 10.53 0.35
Holdings 61°27'40" W (central) Alfredo Gr Devonian
22 Pennzoil & Vict. Julia - 1 20°36'05" S 1971 Carandaity 1.281 m Up. San Upper Oil show 11.52 0.34
Holdings 61°37'03" W (central) Alfredo Gr Devonian
23 Pennzoil & Vict. Katerina - 1 20°44'30" S 1971 Carandaity 1.143 m Up. San Upper — 11.07 0.34
Holdings 61°33'50" W (central) Alfredo Gr Devonian
24 Pennzoil & Vict. Luciana - 1 20°10'40" S 1972 Carandaity 819 m Up. San Upper — 10.26 0.37
Holdings 61°43'10" W (north) Alfredo Gr Devonian
25 Pennzoil & Vict. Marta - 1 20°16'31" S 1972 Carandaity 828 m Up. San Upper — 10.08 0.35
Holdings 61°40'27" W (north) Alfredo Gr Devonian
26 Pennzoil & Vict. Nola - 1 20°07'49" S 1972 Carandaity 760 m Up. San Upper — 9.90 0.36
Holdings 61°47'13" W (north) Alfredo Gr Devonian
27 Pennzoil & Vict. Olga - 1 21°25'13" S 1972 Carandaity 1.172 m Up. San Upper — 11.70 0.39
Holdings 61°52'41" W (central) Alfredo Gr Devonian
28 Pennzoil & Vict. Don Quijote - 1 21°37'47" S 1972 Carandaity 2.895 m Cerro León Lower Oil show 22.68 0.54
Holdings 61°56'43" W (south) Gr Ord./Sil.
29 Repsa & Cía. Palo Santo - 1 23°10'20" S 1974 Pirity 3.763 m Cerro León Silurian — 24.30 0.46
Petrolera del Chaco 60°46'08" W Gr
30 Esso, Aminoil & Berta - 1 22°32'47" S 1976 Pirity 4.789 m Berta Fm Upper — 29.52 0.47
Chaco Expl.Co. 61°00'38" W Jurassic
(continues)
Appendix: CHACO BASIN (PARAGUAY): CHARACTERISTICS OF HYDROCARBON EXPLORATION WELLS (continued)

Geothermal
No. Operator Well Name and No. Location Year Subbasin or Total Depth TD: Geologic Formation Hydrocarbon BHT (°C) Gradient
Area (m) I ndication (°C/100m)
31 Texaco & Marat. Cerro León - 1 19°49' S 1976 Lagerenza high 1.970 m Cerro León Lower Gas show 18.90 0.60
Co. 60°56' W Gr Ord./Sil.
32 Texaco & Marat. Toro - 1 20°07'58" S 1977 Fte. Olimpo 3.418 m Cerro León Lower Oil show 27.27 0.59
Co. 58°57'04" W high Gr Ord./Sil.
33 Texaco & Marat. Gato - 1 20°03'30" S 1978 Fte. Olimpo 1.646 m Cerro León Lower Oil show 15.30 0.50
Co. 58°52'30" W high Gr Ord./Sil.
34 Chaco Expl.Co. Parapiti - 1 21°00'00" S 1977 Carandaity 3.000 m Cerro León Lower Gas show 21.96 0.52
61°00'00" W (east) Gr Ord./Sil.
35 Chaco Expl.Co. Parapiti - 2 21°34'00" S 1977 Carandaity 2.370 m Cerro León Lower — 19.80 0.54
62°00'00" W (central) Gr Ord./Sil.
36 Cía. Petrolera del Anita - 1 22°53'24" S 1978 Pirity 4.129 m Berta Fm Upper — 27.00 0.48
Chaco 61°30'18" W Jurassic
37 Cía. Petrolera del Gloria - 1 22°56'55" S 1979 Pirity 4.016 m Berta Fm Upper — 25.74 0.46
Chaco 60°38'04" W Jurassic
38 Occidental Carmen - 1 23°15'07" S 1985 Pirity 4.511 m Cerro León Silurian — 31.14 0.54
61°18'14" W Gr
39 Occidental Tte. Acosta - 1 22°44'55" S 1987 Pirity 4.268 m Berta Fm Upper — — —
60°25'15" W Jurassic
40 Occidental Nazareth - 1 22°39'17" S 1987 Pirity 4.025 m Berta Fm Upper — 24.93 0.43
59°51'37" W Jurassic
41 Cano Martinez Independencia -1 20° 09' 45"S 1993 Carandaity 609 m Up. San Upper Gas blow — —
61°46' 24"W (north) Alfredo Gr Devonian
Phanerozoic Tectonics and Sedimentation, Chaco Basin of Paraguay, Hydrocarbon Potential
205
Evidence for a Middle–Late Paleozoic Foreland Basin
and Significant Paleolatitudinal Shift, Central Andes
P. E. Isaacson
E. Díaz Martínez
Department of Geology
University of Idaho
Moscow, Idaho, U.S.A.

Abstract

D evonian–Permian data of western Bolivia and adjacent regions are used to construct a paleogeography
of the central Andes. Four phases characterize the sedimentation history. (1) Shallow marine clastic
deposition occurred through the Devonian (Lochkovian–Frasnian), with an increase in sedimentation in
Emsian–Eifelian time. Lithofacies distribution and sediment thicknesses indicate primarily a western source.
Endemic, high-latitude (>55˚ S) fauna with several megafaunal originations in Bolivia also contain organisms
characteristic of North Africa and northeastern United States (Middle Devonian megafaunas and Late
Devonian palynomorphs). (2) Latest Devonian–Early Carboniferous (Famennian–Viséan) sedimentation is
characterized by glaciomarine and fan-deltaic sedimentation. Clasts are derived from underlying sedimentary
units and andesitic, granitic, and tuffaceous rocks. (3) A middle Carboniferous (Serpukhovian–Bashkirian)
hiatus in sedimentation occurred, its age and duration varying across the region. (4) Siliciclastic and carbonate
deposition occurred in Late Carboniferous–middle Permian time (Moscovian?–Artinskian). The clastics were
derived from a western source. Carbonate rocks (Copacabana Formation) were deposited in situ, in low
latitudes (≤ 25˚ S lat). Devonian sedimentation is inferred to have occurred on continental crust in a foreland
setting, with a western magmatic arc. Restoration of the San Nicolás batholiths (U-Pb zircon, 425 and 394–388
Ma) relative to the Devonian basins suggests that they may have constituted the magmatic arc. Intra-arc basins
may have existed near present-day coastal Peru. Following the middle Carboniferous hiatus, sedimentation
continued in a back-arc region, although differentiation of distinct Carboniferous and Permian basins, the
intrusion of plutons inboard of and within the basin along strike, and extensional faulting in the Late Permian
indicate major changes in the tectonic setting, possibly including a reorientation of the subducting slab to a low
angle. Uncertainties in the tectonic setting interpretation are introduced by the incomplete stratigraphic record,
which is obscured in the Altiplano and Cordillera Occidental, and by the undefined history of plate boundary
interactions, such as possible postdepositional strike-slip motion and tectonic erosion along the plate margin.

Resumen

E l estudio de la secuencia del Devónico a Pérmico del oeste de Bolivia y zonas adyacentes permite la
reconstrucción de la paleogeografía de una parte de los Andes Centrales, cuya historia sedimentaria
puede dividirse en cuatro fases. (1) Sedimentación siliciclástica en una cuenca marina somera durante la mayor
parte del Devónico (Lochkoviano a Frasniano), con un aumento de la sedimentación durante el Emsiano y
Eifeliano. La distribución de facies y espesor de sedimentos indican un área fuente situada hacia el oeste. La
fauna endémica de latitudes altas (>55˚ S), con algunos taxones característicos de Bolivia, se mezcla con organ-
ismos característicos del norte de Africa y NE de Estados Unidos (megafaunas en el Devónico medio y pali-
nomorfos en el Dévonico superior). (2) La sedimentación del Devónico tardío y Carbonífero inferior
(Fameniano a Viseano) está caracterizada por depósitos glaciomarinos y de abanicos deltaicos. La composición
de los clastos es de rocas sedimentarias de las unidades subyacentes y de granitoides, andesitas y tobas
volcánicas. (3) Durante el Carbonífero medio (aproximadamente Serpukhoviano y Bashkiriano) tiene lugar
una interrupción en la sedimentación, de edad variable según las zonas, que resulta en la erosión parcial o total
de la secuencia del Carbonífero y Devónico subyacente. (4) Durante el Carbonífero superior y hasta el Pérmico
medio (Moscoviano a Artinskiano) tiene lugar sedimentación mixta siliciclástico-carbonatada. Las areniscas
indican un área fuente al oeste. Las rocas carbonáticas fueron depositadas en latitudes bajas (≤ 25˚ S). La sedi-
mentación del Devónico tuvo lugar en un ambiente tectónico de antepaís situado entre un arco magmático
marginal al oeste y un cratón al este. La posición de los batolitos de San Nicolás en Peru (datado en 425 y
394–388 Ma) en relación a la cuenca sedimentaria sigiere que estes plutónes habrían formado parte del arco
magmático. Los depósitos devónicos del SW de Peru constituirían el relleno de cuencas de intra-arco, mientras
que faltaría el registro sedimentario de la zona proximal de la cuenca de antepaís, probablemente cubierto bajo
el Altiplano y la Cordillera Occidental. La reorganización de la cuenca e intrusión de plutones durante el
Carbonífero y Pérmico, así como la tectónica extensional durante el Pérmico superior, indican reajustes e inter-
relaciones entre los bordes de placas posiblemente relacionados con desplazamientos transcurrentes y dismin-
ución del ángulo de subducción de la placa.

Isaacson, P. E., and E. Díaz Martínez, 1995, Evidence for a middle–late Paleozoic foreland 231
basin and significant paleolatitudinal shift, central Andes, in A. J. Tankard, R. Suárez S.,
and H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 231–249.
232 Isaacson and Díaz Martínez

Figure 1—Geography of central Andes mountains and Figure 2—Simplified tectonic map of Bolivia, with tectonos-
adjacent regions discussed in this paper. tratigraphic domains. 1, Altiplano; 2, Huarina fold and
thrust belt; 3, Cordillera Real; 4, Tarija-Teoponte; 5,
northern sub-Andean fold and thrust belt; 6, central and
southern sub-Andean fold and thrust belts; 7, Brazilian
shield. Thrusts (from Spanish): CANP, main Andean thrust;
INTRODUCTION CCR, Cordillera Real thrust; FCC, Coniri thrust. (After
Sempere, 1987, 1990.)
Several recent syntheses of Devonian stratigraphy,
biostratigraphy, tectonics, and igneous events in Bolivia,
Peru, and northern Chile (Figure 1) permit a tentative
reconstruction of middle Paleozoic events in that region. Huarina fold and thrust belt (Figure 2). The main tectonic
Earlier workers (e.g., Ahlfeld and Branisa, 1960) had structure of this belt is the stacking of Upper Ordovi-
suggested that the thick Devonian clastic rocks in the cian–Lower Triassic sedimentary rocks in a series of
region were deposited in a “geosynclinal” setting. Other tightly imbricated folds and thrusts verging to the
work has supported a passive margin during Devonian southwest. The Huarina belt is separated from the
time (Zeil, 1979), with Late Ordovician and Late Cordillera Real domain to the northeast by the Cordillera
Carboniferous–Permian orogeny bracketing the other- Real thrust and from the Altiplano domain to the
wise quiescent sequence. southwest by the Coniri thrust. Although the sedimen-
The Bolivian Central Andes comprises a series of fault- tary sequence was partially eroded and deformed during
bounded tectonostratigraphic domains (Sempere et al., the Hercynian orogeny of Dalmayrac et al. (1980) and
1988) (see Figure 2), which are substantially different from Martínez (1980), most of the tectonism took place during
previous physiographic divisions used in Bolivian the several Andean (i.e., much younger) phases of defor-
geologic literature (Ahlfeld and Branisa, 1960; Russo, mation.
1966; Rodrigo and Castaños, 1978). These domains It has been suggested that the Paleozoic sedimentary
display a distinct tectonic shortening and relative rocks in the Altiplano of Bolivia are the result of complex
displacement as a result of Cenozoic Andean orogeny filling of a northwest-trending back-arc basin encom-
(Sheffels, 1990; Roeder and Chamberlain, 1995). However, passing most of the present-day central Andes (Sempere,
they are not to be confused with allochthonous terranes 1989). Closure of the basin, together with uplift and
because they have all been linked at least since the early thrust faulting during the Mesozoic and Cenozoic
Paleozoic and are not exotic to South America. With Andean orogenic cycle, resulted in the construction of
respect to these domains, the middle and late Paleozoic of the present Andes Cordillera. The overprint of this defor-
the northern Altiplano is exposed in the northern third of mation complicates the study of the pre-Andean history
the Charasani-Ayoma-Atocha unit, also called the of the western margin of Gondwana.
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 233

Figure 3—Devonian, Carboniferous, and Permian lithostratigraphy, central Andes and adjacent regions. (From Newell et al.,
1953; Díaz, 1959; Eckel, 1959; Dalmayrac, 1977; Isaacson, 1977; Marocco, 1977; Isaacson et al., 1985; Oller and Sempere, in
Duarte, 1989; Bahlburg and Breitkreuz, 1991; Starck et al., 1993.)

STRATIGRAPHY In northern Chile (Atacama Desert region), Middle


Devonian–Lower Carboniferous shallow water conglom-
Devonian Siliciclastic Sequence erates and litharenites contain andesite and diorite clasts
(Davidson et al., 1981; Isaacson et al., 1985; Isaacson and
Basic Bolivian Devonian stratigraphic data essential Sablock, 1988). Also, Upper(?) Devonian turbidites in
for paleogeographic reconstructions are well established northern Chile show a source (Arequipa massif) to the
(Isaacson, 1975, 1977; Oller and Sempere, in Duarte, 1989; north-northwest (Bahlburg, 1985; Breitkreuz and
Starck et al., 1993) and are shown in Figure 3. The bulk of Bahlburg, 1985; Bahlburg et al., 1986).
the thick clastic sequence is primarily late Early (Emsian) Source areas (Figures 4, 5) for the Bolivian sequence
and early Middle Devonian (Eifelian) in age, although are the ancestral Arequipa massif (Isaacson, 1975; Boucot
newer information from ammonites (Babin et al., 1991) et al., 1980; Laubacher et al., 1982), the Puna highland of
and palynomorphs ( Le Hérissé et al., 1992; Racheboeuf northwestern Argentina, and probably the Brazilian
et al., 1992) suggests that Givetian, Frasnian(?), and shield (Isaacson and Sablock, 1988, 1990). These areas
Famennian rocks are also present. Since there are few supplied sediments to northwestern-central, southern,
megafossils, palynomorphs in the upper part of the and eastern Bolivia, respectively. Principal constituents
sequence suggest presence of the Famennian–Tour- of the strata appear to reflect compositional differences in
naisian boundary (Lobo-Boneta, 1987; Vavrdová et al., these source terranes. In the Altiplano of Bolivia, mature,
1991). Frasnian units are present in the sub-Andean and fine-grained Upper Devonian sandstones suggest mixed
subsurface regions of eastern Bolivia and Peru (Barrett sources from both the craton and a magmatic arc, with
and Isaacson, 1988; Isaacson and Sablock, 1990). Primary little reworking.
lithologies in the western and central Bolivian sequence The Arequipa massif was largely sialic (Bellido and
are (in decreasing abundance) siltstone, quartz arenite, Guevara, 1963; Cobbing et al., 1977; Dalmayrac et al.,
mudstone, and shale (Isaacson and Sablock, 1988). Felds- 1980) and may have been part of Miller’s (1970) “Pacific
pathic sandstones are found in southern Bolivia. Lithofa- continent.” The Puna highland is composed of Upper
cies and isopach modeling of the Devonian indicates Ordovician magmatic igneous rocks (Davidson et al.,
coarser and much thicker lithologies in northwestern 1981; Coira et al., 1982; Palma et al., 1987). (These terranes
Bolivia than those in eastern and southern Bolivia are discussed further later.) The Puna also influenced
(Isaacson, 1975). Atacama (Chilean) depositional events, including
234 Isaacson and Díaz Martínez

Figure 4—Paleogeographic map of the central Andes for Figure 5—Paleogeographic map of the central Andes for
late Early Devonian (Emsian) time. Major structural Middle Devonian time (Eifelian–Givetian). Maximum trans-
elements responsible for the Bolivian-Peruvian basins gression, subsidence, and deposition is within the basins
include the Late Ordovician accretionary volcanic- at this time, as shown by sedimentation within the
magmatic Puna terrane, sialic Arequipa massif (with its Solimoes and Arce basins (da Silva, 1988). The Madre de
Early Devonian San Nicolás batholiths), and Brazilian Dios basin also received significant sedimentation at this
shield. (Latitudes from Isaacson and Sablock, 1990.) The time (Isaacson et al., 1995).
“Bolivia basin” (Frutos and Tobar, 1975; Suárez-Soruco
and Lobo-Boneta, 1983) includes the Beni basin (Díaz,
1959), whose margins have yet to be defined. The “Peru Late Devonian–Lower Carboniferous
basin” (Mosmann et al., 1986, his figure 11) includes the Diamictites and Related Rocks
Ucayali and Madre de Dios basins, the latter possibly
having limited stratigraphic development in the Early Located above shallow marine shales (Colpacucho
Devonian (Isaacson et al., 1995). The northwest Peru Formation) and at the base of the Cumaná Formation in
orogeny, proposed by Isaacson (1989), appears as a
the northern Altiplano of Bolivia is a diamictite unit
highland in the reconstruction of deMelo (1988).
(Figure 6), which can be traced for more than 30 km
along strike from Isla del Sol to Cumaná (Lake Titicaca
area), with a variable thickness of 60–70 m (Díaz and
Lema, 1991a). It is dated as Famennian (pusillites–lepido-
westward-propagating, shallow water clastic deposits phyta palynozone) by Vavrdová et al. (1991, 1993), and it
with a nearby magmatic terrane source (Davidson et al., may continue into the Tournaisian. Clasts within the
1981; Isaacson et al., 1985). The Brazilian shield is compo- diamictite are subangular to well rounded and range
sitionally similar to the Arequipa massif (Dalmayrac et from coarse sand to boulder size. Clasts are composed of
al., 1980). Orogeny folded and metamorphosed quartz arenite, granitoid, quartzite, conglomerate, and
Devonian and lower Paleozoic flysch in central Peru, intermediate volcanic rock (Díaz et al., 1993). Their
although these effects are not evident in the Bolivian and variable composition, together with some striated and
northern Chilean Devonian sequences. Also, part of the faceted clasts, suggests a glaciated heterogeneous source
Devonian sequence may have been eroded in central terrane.
Peru, and Lower Carboniferous (Tournaisian) clastic There are two major lithofacies associations. The
rocks unconformably overlie it (Mégard, 1973). Evidence lower is dominated by laminated mudstone with drop-
of orogeny decreases away (east-southeastward) from stones, interpreted as ice-rafted and suspended sediment
central Peru, such that only a slight angularity exists deposits. Deposited next on an erosional base is a
between Devonian and Carboniferous rocks in Bolivia massive, matrix-supported diamictite with deformed
(Laubacher, 1974, 1977; Isaacson, 1975). sandstone lenses and boulders up to 2 m long. This is
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 235

interpreted as the result of subglacial deposition at the


margin of a partially floating ice mass, with resedimenta-
tion by postdepositional mass movement processes
(subaqueous slumping), along with partial reworking by
currents. This diamictite may be the earliest sign of late
Paleozoic glaciation in South America. Paleogeographic
reconstruction of the region during the Early Carbonif-
erous suggests that the Eohercynian orogen in the
Cordillera Real is a possible source terrane (Dalmayrac et
al., 1980). Lower–middle Carboniferous clastic units of
this area (Cumaná and Kasa formations) (Figure 7) show
continued glaciation of the source terrane through much
of this sequence.
The principal glacial episode affecting Gondwana
during the late Paleozoic took place from Serpukhovian
(Namurian A) to Sakmarian, with two minor episodes in
the Famennian and Viséan (Veevers and Powell, 1987).
Evidence for the two short episodes comes from marine
and nonmarine deposits in Brazil and northwestern
Africa, as well as recent work in Bolivia (Vavrdová et al.,
1991, 1993).
The base of the Cumaná Formation at Isla del Sol has
yielded abundant specimens of organic microfossils,
including Retispora lepidophyta and Umbellasphaeridium
saharicum (Vavrdová et al. 1991, 1993). These occurrences,
along with Hymenozonotriletes explanatus (and the absence
of Verrucosisporites nitidus) gives a late Famennian (Fa 2c-d),
palynozone PL (pusillites-lepidophyta) age for the diamictite.
Because original samples were taken on a reconnaissance
basis, incomplete sampling at this locality has not
permitted an upper age bracket for the diamictite. Dating
of equivalent strata at Mina Matilde (about 40 km north
of Isla del Sol on the east shore of Lake Titicaca), where
the diamictite does not exist, suggests that the Famen-
nian and Tournaisian transition is present.

Carboniferous Siliciclastic Units


Carboniferous deposits in the Altiplano accumulated
in a northwest-southeast trending basin. Mountain ranges
east and west of the basin supplied sediment during part
of the Carboniferous. The Carboniferous includes
deposits from the Ambo and Titikaka groups. The
Cumaná, Kasa, and Siripaca formations (Figures 3, 7)
record accumulation in the late Famennian–late Viséan
(Rösler et al., 1989), and the Yaurichambi and Copaca-
bana formations record accumulation in theKasimovian–
Artinskian (Sakagami et al., 1986; Sempere, 1987, 1990).
The Ambo Group of Peru is associated with regional
climatic change and tectonic instability in the areas
providing sediment to the Carboniferous basin. These
deposits, together with glacial deposits in the Solimöes,
Amazonas, and Parnaiba basins of Brazil, represent a
major period of glaciation over much of Gondwana
(Veevers and Powell, 1987). In the Altiplano area,

Figure 6—Detailed lithologies of the Cumaná Formation at


Isla del Sol. Famennian deposition represents a restricted
shelf with increasing ice-rafted lithoclasts. Thickness in
meters; black areas represent mudstone.
236 Isaacson and Díaz Martínez

Figure 7—Vertical profiles typical of Carboniferous deposits in the northern Altiplano of Bolivia. Sections: 1, Copacabana
area (Hinchaka, Santa Ana, Siripaca, Belén); 2, Cumaná; 3, Calamarca; 4, Carabuco; 5, Ancoraimes; 6, Yaurichambi; 7,
Colquencha. Dual numbers indicate formations: 1.2, Copacabana Formation (also Permian in part; see Figure 11); 2.1,
Yaurichambi Formation; 1.3, Siripaca Formation; 1.2, Kasa Formation; 1.1, Cumaná Formation (see Figure 10). D, Devonian.
Lithologic symbols: a, limestone (with dolomite and sandstone); b, mudstone, including coal (Siripaca Formation); c,
sandstone, including conglomerate; d, diamictite, conglomerate, and sandstone.

adjacent ranges developed an ice cover that existed It is not necessary to propose changes in climate,
during the late Famennian, Tournaisian, and part of the tectonics, or source area to explain the differences in
Viséan. The scattered diamictites that accumulated in a modal composition and textural characteristics between
fan delta setting (Kasa Formation) record the gradual the fine-grained, quartz-rich sandstones of the lower
retreat of the glaciers to the ranges. Increase of sediment member of Kasa Formation and the feldspathic sand-
input and subsidence rates possibly related to activity in stones of its upper member. It has been shown that
the magmatic arc also influenced accumulation patterns hydraulic sorting, current reworking, and differential
in the Kasa Formation. The locally unconformable weathering processes present in coastal areas and
character of the base of this cycle is thus most likely shallow clastic shelves can result in such differences
related to the erosional and depositional processes (Mack, 1984; Dickinson, 1988). These processes are
typical of a glaciomarine environment, instead of enhanced by storm and wave reworking. Suttner et al.
emersion and subaerial erosion. Apart from the erosive (1981) presented some interesting examples of modifica-
character of the gravity flows and mass transports tion of petrofacies due to different sedimentary environ-
common in a glaciomarine environment, it is possible to ments, comparing alluvial fan and clastic shelf deposits.
develop erosional discontinuities in deep marine envi- They concluded that the same provenance area but a
ronments during glacial periods as a result of bottom different transport distance and depositional environ-
density currents from adjoining glaciated areas (Johnson, ment can result in different petrofacies. This case study is
1974). similar to what is found in the Kasa Formation.
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 237

The depositional setting and modal composition of graphy, producing deposits with variable thicknesses on
the coarse feldspathic sandstones in the Kasa Formation the east shore of Lake Titicaca to the north and in the
were greatly affected by the cold climate and sparse Colquencha zone, 60 km south of La Paz, in the Altiplano.
vegetation during their deposition as evidenced by the Paleosols in the Yaurichambi Formation indicate a warm
few plant fossils found in it. Azcuy and Suárez-Soruco climate during this part of basin evolution, continuing
(1984) mention equisetae, lycopsidae, and monopinnate with accumulation of the carbonate deposits in the
plant fronds from the Kasa Formation, which they Copacabana Formation.
assigned to Nothorhacopteris. This genus has been found As shown in Figure 8, the Copacabana Formation in
in the Kaka Formation (Figure 3), which has an Early northwestern Bolivia shows a complex interplay of sand-
Carboniferous palynologic assemblage, a recycled Late stones, terrigenous mudstones, various limestones,
Devonian assemblage, and a total absence of gymno- dolomites, caliches, and other lithologies (Isaacson et al.,
sperm pollen. 1993). Lithofacies change significantly over short
The transition of the Kasa to the Siripaca Formation, distances, both up-section and in a west-east direction.
however, marks an important change in the paleogeo- For simplicity, we have divided the formation into two
graphy and paleoclimate of the basin, including the principal units: the lower unit, comprising cycles 1 and 2,
disappearance of glaciers as temperatures rose, an and the upper unit, comprising cycles 3 and 4 (Figure 8).
increase in vegetation cover, and a reduction of relief in The lower unit tends to be dominated by siliciclastic
the source areas. Climatic change resulted from rocks, whereas the upper unit is mainly carbonates.
movement of this part of Gondwana to lower latitudes Beginning at a moderate-relief (30–40 cm) caliche
(Caputo and Crowell, 1985; Veevers and Powell, 1987; surface (at Yampupata and Yaurichambi), fine-grained
Díaz et al., 1993). During this same period, however, the sandstone was deposited, followed by greenish gray
global paleoclimate picture changed to one of general (tuffaceous) siltstone with thin (2–3 m) beds of extensively
cooling (Veevers and Powell, 1987; Sablock, 1993). bioturbated carbonate mudstone. Next are sandstone
channels, with more persistent bioturbated mudstones
Middle Carboniferous Hiatus (containing large productoid brachiopods) and green silt-
stones following them. Another caliche occurs in
The Siripaca Formation represents the filling of the siltstone–sandstone, followed by lime mudstone–wacke-
Bolivia basin and the final Paleozoic regression to the stone with fenestral fabric and tepee structures. An exten-
north or northwest. At the Copcabana Peninsula, this sively cross-bedded, coarse-grained sandstone (shoal?)
formation has a small (1.5-m-thick) coal bed. Its age and occurring above another caliche terminates the lower
rank are not known, although it has been excavated for lithofacies. At Ancoraimes (northeast of the Peninsula de
local use. During the Serpukhovian and Bashkirian, a Copacabana sections), initial results suggest less obvious
period of intense subaerial erosion and nondeposition caliches, as well as alternating thinner sandstone and
followed in the Altiplano, while deposition continued in wackestone–packstone carbonates. Separating the two
adjacent basins of the central and southern sub-Andean basic lithofacies is another caliche, which is followed by
(Macharetí Group) and northern sub-Andean regions lenticular medium-grained sandstone, mudstone, and
(Kaka Formation (Sempere, 1990). As a result of the shell-rich laminated sandy dolomite, suggestive of
northward retreat, the stratigraphic gap recorded by the tempestites. Green, laminated, volcanogenic (Isaacson et
middle Carboniferous unconformity is quite variable. al., 1995) mudstone follows, punctuated by well-biotur-
This final regression is probably related to two global bated and fossiliferous mudstone–wackestone. Green,
events that took place during the Namurian A or late cross-bedded, coarse-grained sandstone (including some
Viséan to Serpukhovian (Saunders and Ramsbottom, litharenite) interbedded with claystone and minor
1986), which were produced by the onset of the main mudstone (with current-stable brachiopod shells)
glacial episode in Gondwana (Veevers and Powell, 1987). comprise the uppermost beds of this unit. Several caliches
also occur. A significant bioturbated and fossiliferous
Upper Carboniferous–Lower Permian packstone–grainstone caps the sequence at Yampupata,
Carbonates and Siliciclastics where the uppermost beds may be missing.
Units overlying the Copacabana Formation are
Marine deposition resumed in the Kasimovian (or variable. Where significant erosion is not apparent,
even earlier in the Bashkirian; Isaacson et al., 1995) when presumably by fluvial systems depositing sandstones,
transgression represented by the Yaurichambi and there is a gradual changeover to siliciclastic-dominated
Copacabana formations occurred. Transgression and minor carbonate lithologies (such as at Cumaná). A
proceeded slowly, as seas advanced from the north. complex suite of lithologies terminates the Copacabana
Progressive changes in local base level produced a Formation.
blanket of fluvial, deltaic, and coastal deposits Using several features that indicate subaerial exposure
(Yaurichambi Formation) at the base of the transgressive as defined by Flügel (1982) and Esteban and Klappa
sequence. Paleocurrent orientations are from several (1983), eustatic cycles within the Copacabana Formation
directions in this sequence, recording a major reorganiza- are apparent. The features include fenestral fabric, tepee
tion of basin drainage system (Sempere et al., 1986; structures, and caliches. Complicating the cycles,
Marocco et al., 1987; Barrios and Beccar, 1988). The trans- however, are apparent tectonic adjustments of the basin.
gressive deposits filled the newly created paleotopo- For example, cycles 1, 2, and 4 (Figure 8) have terrige-
238 Isaacson and Díaz Martínez

Figure 8—Lithostratigraphy, biostratigraphy, and previous stratigraphic nomenclature of the Copacabana Formation (Late
Carboniferous–Early Permian), western Bolivia. Cycles, hiatuses, caliches (triangles), and sedimentary petrology are as
discussed in this paper and in Isaacson et al. (1993).

nous influxes from the west that may have abruptly d’Orbigny, 1842), as well as unusual in abundance and
infilled the basin, such that they substituted shallower diversity, has also been thought to be part of a cold-
depositional systems for slightly deeper ones. Cycle 3 water “austral” province (Clarke, 1913). Late Carbonif-
contains lagoonal claystone, which indicates temporary erous–Permian faunas have also been described
quiescence in the western source area(s). We suggest that (Dunbar and Newell, 1946; Newell et al., 1953) and have
even with Early Permian marine onlap (Vail et al., 1977), been biogeographically linked to West Texas in the
the northwestern Bolivian depositional basin maintained United States.
a position close to sea level.
Devonian Paleobiogeography
PALEOBIOGEOGRAPHY Devonian paleobiogeography of the central Andes
was influenced by three major factors: the paleogeo-
Work on the paleobiogeographic significance of the graphic setting, the high latitudinal position (and cold
central Andean Devonian–Permian faunas has depended temperatures) of the region during Early Devonian time
on the abundance and components of recovered faunas. (Figure 4), and the apparent influx of slightly warmer
The Devonian, long considered to be “classic” (e.g., water during Middle Devonian time (Figure 5). The
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 239

Figure 9—Paleobiogeographic map of the central Andes Figure 10—Paleobiogeographic map of the central Andes
for Emsian time. High-latitude cold water currents main- for Middle Devonian time. Lower latitudes (see Figure 15)
tained the low-diversity Malvinokaffric faunas. Vestigial and transgression permitted entry of Tropidoleptus and
warm water currents from Colombia and possibly interior other taxa from North Africa. The residual Malvinokaffric
North America are responsible for the Eastern Americas stock supported a limited diversity in Bolivia and adjacent
Realm faunas of the possible intra-arc basins of southern regions. (From Isaacson and Sablock, 1990.)
Peru. Connections to the Amazonas basins are ephemeral.
(From Isaacson and Sablock, 1990.)
Early–Middle Devonian time was probably influenced
by extensive basin development within Peru, Chile, and
highly endemic Malvinokaffric faunas (Richter and Bolivia. The high latitudinal position of this region
Richter, 1942) of southern South America, Antarctica, allowed colder water Malvinokaffric fauna to colonize
and South Africa have long been assumed to have lived much of South America. The lowest latitudinal penetra-
in cold water (Figure 9), with evolutionary ties to Eastern tion of Malvinokaffric faunas (e.g., Australocoelia, as
Americas Realm organisms of the northern Appalachian reported by Mégard, 1973) roughly corresponds to a
basin in the United States (Boucot, 1971; Isaacson, 1977). postulated cold water subpolar gyre that paralleled the
It has been suggested that Early Devonian Eastern western margin of South America up to about 40˚ S lat
Americas Realm brachiopods in interior southern Peru (Figure 9). During Middle Devonian time, subsequent
(Boucot et al., 1980) entered the region by means of an northward movement of western Gondwana coupled
influx of warmer water and that the migration routes with marine transgression introduced the warmer water
may have been restricted by intra-arc basins (see Figure Eastern Americas Realm fauna farther south into Bolivia.
14). Higher in the Andean sequence, however, the much Also, selected brachiopod taxa arrived from North
lower diversity post-Malvinokaffric fauna (Figure 10) has Africa. Isaacson and Perry (1977) suggest that Tropi-
been tentatively identified Isaacson and Sablock, 1990). It doleptus arrived in Bolivia via the Amazonas basin. Other
consists of the circum-Atlantic brachiopod genus Tropi- taxa (e.g., Globithyris or Rhipidothyris) may also have
doleptus (Isaacson and Perry, 1977) and other taxa. It taken advantage of this seaway, although the full
appears that Globithyris (or Rhipidothyris), which is above complement of North African taxa is not found in
Tropidoleptus in the Devonian sequence (Isaacson, 1974), Bolivia.
may have moved from Libya, where both taxa are
present in the Idrí Formation of Givetian age (Havlicek Carboniferous Paleobiogeography
and Röhlich, 1987). This Middle Devonian fauna,
however, does not achieve the diversity of Hamilton and While there are few Bolivian Carboniferous marine
other Givetian fauna in New York and North Africa. megafaunas, Isaacson et al. (1985) reported on Early
Distinct faunal provinciality exhibited during Carboniferous (Tournaisian) brachiopods from northern
240 Isaacson and Díaz Martínez

Chile (see Figure 14). Dutro and Isaacson (1991) have (1981), and it is quite distinct from the nearby
shown that this assemblage is characterized by relatively Durhaminid Province in the western and northwestern
large, stout shells in several brachiopod families, United States. Considering the biogeographic separa-
although it is generally of low diversity. These “big tions within North America, it is remarkable that a single
shells” also occur in central Peru, northwestern Argen- South American–North American province existed,
tina, eastern México (Oaxaca and Tamaulipas), eastern thereby demonstrating a close paleogeographic link
Appalachians (eastern West Virginia), Libya, Iran, and between the two regions.
southeastern Australia, suggesting that the Chilean Ross (1973) defined a mid-Andean and mid-continent
occurrence represents a relatively cosmopolitan (albeit (southern United States) biogeographic unit that extends
facies-controlled) fauna. through Mexico and Morocco to south of Santiago, Chile.
Rösler et al. (1989) have described a “pre-Glossopteris Ross further stated that the Mid-Continent (U.S.)
flora” of Early Carboniferous age from the Siripaca Fusulinid Province of the Late Carboniferous–Early
Formation. They suggest, furthermore, that the plant Permian was paralleled by rugose corals and ammonoid
fossils, including Nothorhacopteris, Paracalamites, and lyco- cephalopods. Gobbett (1973) identified distinct links with
phytes, show little transport and were deposited in Tethyan forms, as well as widespread cosmopolitan taxa.
nearshore sediments of the Siripaca Formation. Erwin et Boucot and Gray (1979) described a southern (warm)
al. (1992) have identified a warm, temperate Carbonif- marine surface circulation gyre, which equally affected
erous flora from coastal Peru, which was part of a coastal Gondwana (South America) and the southern
climatic belt between the tropical Euramerican Realm United States, while a northern gyre affecting western
and the cool Gondwana Realm in Bolivia. North America allowed contrasting faunas to exist over a
Newell et al. (1953) have discussed the significance of short paleogeographic distance.
Pennsylvanian marine faunas from Peru. They are
cosmopolitan, although they have a slightly reduced
diversity (compared to coeval North American faunas).
Until more palynomorph and paleobotanical informa- PALEOGEOGRAPHY
tion is forthcoming from Bolivia, however, little can be
concluded about its biogeographic placements. Early–Middle Devonian Paleogeography

Permian Paleobiogeography Current Devonian paleogeographic reconstructions of


the western continental margin of Gondwana in Peru
Following the work of Dunbar and Newell (1946) on and Chile (Figures 4, 5) demonstrate that sedimentation
fusulinids, the Copacabana Formation carbonate rocks on this margin was influenced by at least two different
are considered to be primarily Wolfcampian. Detailed tectonic sources. The first, the Arequipa massif, with a
fusulinid and conodont biostratigraphic work by superimposed Early Devonian arc (Mukasa and Henry,
Sakagami et al. (1986) and Suárez-Riglos et al. (1987) 1990), supplied a large volume of clastic sediment to the
have confirmed that the carbonate rocks of northwestern Peruvian and Bolivian basins (Isaacson, 1975; Zeil, 1979;
Bolivia are Virgilian (Late Carboniferous) to early Dalmayrac et al., 1980). The second, which is a western
Leonardian (Permian) in age. Newer work (Isaacson et Argentine magmatic arc terrane of Late Ordovician age
al., 1995) suggests that much of the carbonate sequence in (Coira et al., 1982; Mpodozis and Forsythe, 1983; Palma
the northern Bolivian subsurface is even older et al., 1987), influenced northern Chilean and Bolivian
(Morrowan). Major unconformities, while relatively depositional events. Preliminary assessment of the
minor in the Lake Titicaca region, exist between the Devonian–Early Carboniferous paleogeography of this
Virgilian and older strata, as well as within the Leonar- complex region suggests that the Arequipa massif
dian and Guadalupian parts of the section (Collasuyo extended beyond the present west coast of South
Member; Figure 3). (The significance of smaller discon- America and had a southern boundary in northern Chile
formities and paraconformities within the Wolfcampian (Isaacson, 1975; Isaacson and Sablock, 1988, 1990).
strata at Lake Titicaca was described earlier.) The Late Between northern Chile and central Peru is the
Carboniferous unconformity becomes more pronounced Bolivia basin, in which epeirogeny or magmatic-related
(and visible) to the south and east (Barth, 1972; Azcuy et tectonics may have influenced regional sea level fluctua-
al., 1982). tions and unconformities (sometimes angular) between
Micro- and megafaunas of the Permian System in Devonian and Carboniferous strata. There also is a
Bolivia have long been considered as “West Texas” in suggestion that fold axes in Peru and Bolivia indicate
affinity (Dunbar and Newell, 1946). Based on fusulinid that the compressional forces during Late(?) Devonian
occurrences, Newell et al. (1953) defined a Peruvian time were acting in a southwest-northeast direction
fauna of limited distribution, although most taxa (Mégard, 1973; Dalmayrac et al., 1980). The coastal
occurred simultaneously in North America (and other Chañaral melange of northern Chile indicates compres-
cosmopolitan localities) and Andean South America. sion (Bell, 1987). Some evidence indicates that subduc-
Wilson (1990) offered a well-documented correlation of tion began in the north (Peru) and propagated south
similar rugose coral taxa among Bolivia, Peru, (Bahlburg et al., 1986). In central Chile, this new conver-
Guatemala, and the southwestern United States. This has gence emplaced various island arcs and microcontinents
been identified as the Cyathaxonid Province by Hill (Ramos et al., 1986; Bell, 1987).
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 241

Paleogeographically, these events would have paleogeographic reconstructions of Gondwana for the
restricted formerly “open” seaways along eastern Peru, Late Devonian, with the northern Altiplano placed at
western Bolivia, and northern Argentina and Chile 55˚–65˚ S lat (Isaacson and Sablock, 1988, 1990).
sometime during the Devonian. Therefore, Late The offshore marine shelf setting of the units under-
Devonian glaciation in Bolivia (Vavrdová et al., 1991; lying and overlying the upper Colpacucho Formation
Díaz et al., 1992) and along the upper Amazon basin implies a much higher accumulation rate taking place at
(Rocha Campos, 1983; deMelo, 1988) may have occurred the same time in proximal areas of the basin, closer to the
on substantially uplifted “orogenic” terranes in central shore (Powell, 1988). Absence of corresponding facies
and western Peru. However, there were marine seaways and thicknesses at the Carboniferous sections to the east
adjacent to these features and black shale deposition in favor a western or, less likely, southern source for the
the upper Amazon region (Hünicken et al., 1988). diamictites. The isotopic dating and geochemistry of
these granitoid boulders should give some insight into
Late Devonian–Early Carboniferous their provenance and the paleogeography of the basin.
Paleogeography Based on a recently revised lithostratigraphic subdivi-
sion of the Carboniferous system in the Altiplano basin
Recognition of a probable source for the clasts in the of Bolivia (Díaz, 1991; Díaz and Lema, 1991b), a general
diamictite (Cumaná Formation) is of major significance outline of its paleogeographic and paleoclimatic
for the paleogeographic reconstruction of the Carbonif- evolution is proposed. Deposition began with glacioma-
erous basin. The Paleozoic northwest-southeast trending rine deposits (Cumaná Formation) during late
back-arc basin of the central Andes (Sempere, 1989), Famennian time, marking the base of a complex thick-
which was limited on the west by the Precambrian ening and coarsening sequence resulting from coastal
Arequipa massif and on the east by the Brazilian shield, progradation (Ambo Group). This sequence includes
was partially disrupted by the Eohercynian belt in wave- and storm-dominated shallow clastic shelf
Famennian–Viséan time (Dalmayrac et al., 1980; deposits (lower member of Kasa Formation). Interbed-
Sempere, 1989, 1990). Provenance studies of the granitoid ding of these marine units with braided alluvial plain
clasts in the diamictite provide a candidate for the deposits records the progradation of a fluvial- and wave-
glaciated highlands. Large clast size (up to 2 m in dominated fan delta complex (upper member of Kasa
diameter in the Quebrada de Chamacani) and geologic Formation), probably as a result of tectonic and glacial
setting could have preserved the original geochemical activity in local uplifts. The overlying sequence marks an
and isotopic composition in the cores of the boulders. important change in fluvial style, from braided streams
However, there are no known studies dealing with this in the Kasa Formation to meandering streams in the
problem, thus leaving three potential source areas: (1) the Siripaca Formation. This latter unit accumulated in a
Eohercynian belt, limiting the basin at the time to the east fluvially dominated deltaic plain that records the filling
(Sempere, 1987); (2) the early Paleozoic Puna magmatic of the basin under conditions of less sediment influx and
arc (“Faja eruptiva de la Puna”) to the south (Coira et al., lower depositional slope. The Siripaca Formation marks
1982; Allmendinger et al., 1983); or (3) the Arequipa the end of the Ambo Group, followed by an erosional
massif to the west. We suggest that the extensive period that removed Lower Carboniferous and Upper
transport of debris from the Puna would not have Devonian deposits in some areas.
permitted occurrence of the large clasts.
Famennian glaciomarine units in northwestern Late Carboniferous–Early Permian
Bolivia support the concept of a Late Devonian glacial Paleogeography
episode in South America, described by Caputo (1985),
Caputo and Crowell (1985), and Veevers and Powell During the Late Carboniferous, a transgression from
(1987). A single alpine glacial advance and retreat may be the north began, accompanied by migration of alluvial
inferred from the overall sequence and relatively small and coastal deposits into the Altiplano (Yaurichambi
thickness of the unit. Recent studies of both modern and Formation). Development of a shallow carbonate ramp
ancient glaciomarine settings allow for the differentiation followed, which persisted into the Permian (Copacabana
of polar, subpolar, and temperate styles of deposition Formation).
(Crowell, 1988; Matsch and Ojakangas, 1988; Brod- The evolution of the Bolivia basin records important
zikowsky and Van Loon, 1991). Rounded clasts, climatic changes related to the shifting of this part of
suspended sediment plumes and debris flows (indi- Gondwana toward lower latitudes. Contemporaneous
cating episodic sediment influx), abundant dropstones events included global climatic and eustatic changes
and striated clasts, and subglacial meltwater outflow (onset of the main Gondwana glaciation) and regional
deposits overlying the diamictite all imply a warm-based tectonism (Eohercynian deformation). Interaction of
(wet) ice body and a probable temperate glacial setting. these factors influenced Carboniferous sedimentation in
Despite the scarcity of ice-rafted sediment in the modern the Altiplano.
model of temperate glacial deposition from the Gulf of A warm climate apparently occurred in northern and
Alaska (Molnia, 1983, 1988), a high iceberg sedimentation western Bolivia (Yaurichambi and Copacabana forma-
rate is inferred from iceberg sedimentation theoretical tions) at the same time that sedimentation in Argentine
models for temperate glacial environments (Dowdes- and southern Bolivian basins was under glacial influence
well, 1988). This temperate setting is in agreement with (Hambrey and Harland, 1981). This can be explained by
242 Isaacson and Díaz Martínez

Figure 12—Summary of Precambrian and Paleozoic


igneous and metamorphic events that may affect paleo-
geographic development of central Andean Paleozoic
basins. Legend: A, margin of Paleozoic basin; B, Brazilian
shield (>1400 Ma); C, plutons (Ordovician, Devonian, and
late Paleozoic). Brazilian shield, San Ignacio belt, and
Rondonia-Sunsas rocks are described in Litherland et al.
(1985). 1, San Ignacio schist and pluton belt (about 1300
Ma); 2, Rondonia-Sunsas granites (1000 Ma); 3, San
Nicolás batholiths (425, 394, and 388 Ma) (Mukasa and
Henry, 1990); 4, Arequipa massif (>1400 Ma) (Shackleton et
Figure 11—Suggested paleogeography of Permian basins al., 1979); 5, Pampean Ranges, plutons and metamorphic
of Peru, Bolivia, and northern Chile (Marocco, 1977; rocks (475, 440, and 340 Ma) (Rapela et al., 1982); northern
Rodrigo and Castaños, 1978; Rivano and Sepúlveda, 1983). Chilean information is from Boric et al. (1990); 6, Lila
Isopachs in meters. Northern Bolivian (subsurface) infor- plutons (450 Ma); 7, pre-Devonian metamorphic rocks; 8,
mation is incomplete. Paleolatitudinal position from late Paleozoic plutons (318 and 225 Ma); 9, Vilcabamba
Veevers (1984) and Rapalini and Vilas (1991). Arequipa (330? Ma) and Late Permian intrusions (Carlier et al., 1982);
massif (Isaacson and Sablock, 1990) reactivated during 10, Ampares massif (330 Ma) (Carlier et al., 1982).
Permian time (Newell et al., 1953) to provide source for sili-
ciclastics (Isaacson et al., 1993).

the high-latitudinal gradient of temperatures taking ates and evaporites appeared in Pennsylvanian time.
place during the glaciation episodes, creating a warming Given the recent work in northern Chile (e.g., Hervé et
and narrowing of the equatorial zones (Raymond et al., al., 1981; Rivano and Sepúlveda, 1983), however, it
1989; Raymond, 1990; Sablock, 1993). appears that marine connections existed between that
Figure 11 presents a suggested basin configuration for region and northern Bolivia, at least ephemerally. From
Bolivia and northern Chile during Early Permian time. the distribution of Permian rocks, it also appears that the
Isopach information comes from the present study and Potosi salient (Figure 11) separated the main Bolivian
from Rodrigo and Castaños (1978). More recent Bolivian basins from a marine connection to northern Chile.
information (Isaacson et al., 1995), however, suggests that Newell et al. (1953) described a Permian orogeny in the
the Upper Carboniferous–Permian extends into the region and suggested that it caused uplift in the
northern sub-Andean subsurface, with a possible Cordillera Occidental. The admixture of volcanogenic
highland separating the Lake Titicaca (exposed) units detritus with carbonates in northern Chile (Niemeyer et
from areas to the north and northeast. Also, B. Mamet al., 1985) may confirm this suggestion. Newell et al.
(personal communication, 1993) has identified (1953) suggested, furthermore, that the orogeny formed a
“Morrowan” formaminifera from the Pando X-1 and “borderland” along the western coast of South America
Manuripi X-1 wells, thereby documenting that carbon- (Newell et al., 1953). We ascribe this event to the wide-
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 243

spread Gondwanan events of Dalmayrac et al. (1980), Ocona phyllonite zone is a major thrust zone repre-
although sedimentation related to it (Collasuyo senting the northeastern margin of an orogenic belt.
Formation and equivalents) is beyond the scope of this There was a separation of staurolite-andalusite schists
study. (Ocono) from Mollendo granulites; evidence suggests a
The Early Permian was a time of relatively reduced contact between them. The age of the Mollendo gran-
tectonic activity, as the northern margin of Gondwana (at ulites is from Rb-Sr whole-rock work by Cobbing et al.
about 0˚–5˚ S lat at this time) (Rapalini and Vilas, 1991) (1977) and from U-Pb concordia plots. Gneisses (Rb-Sr)
had completed coupling with Laurussia (Ziegler, 1989). near Arequipa give an age of 991 ± 52 Ma. Zircons give a
Assuming that little north-south austral shortening has U-Pb age 1100 ± 50 Ma from granulites (Dalmayrac,
taken place in the latitudes between northern Gondwana 1977). Late Precambrian–early Paleozoic Atico plutonism
and Bolivia–northern Chile (at about 20˚–25˚ S lat; includes granulite facies metamorphic rocks, as well as
Veevers, 1984), we propose that central Chile (Santiago pink granites, pegmatites, aplites, dolerites, gabbros, and
region) would have been at about 40˚–45˚ S lat. The diorites. Deformation occurred about 450 Ma, and a lead
glacial episode in northern Bolivia indicates that the loss occurred during late Precambrian orogenesis, meta-
glacial advances took place before Late Carboniferous morphism, and uplift (600 Ma). There is an early
time. Crowell (1978) suggested that the glacial activity Paleozoic Marcona sequence of low grade metasedimen-
across Gondwana (Paraná basin and Malvinas Islands) tary rocks (greenschist facies). Deposition of these rocks
ended during the latest Carboniferous. It appears that the occurred between 440 and 390 Ma. The San Nicolás
continuing northward drift of Gondwana into lower batholith occurred about 390 Ma (Mukasa and Henry,
latitudes (Rapalini and Vilas, 1991), occurring with the 1990).
initial assembly of Pangea, contributed to the demise of
glaciation. Onshore prevailing winds at low southern Oriente of Peru and Brazilian Shield (Bolivia)
latitudes transported across nearly 8000 km of land
(eastern and central Gondwana) would have been mois- In Peru, there were three major orogenic cycles
ture deficient, producing an arid zone across western (Carlier et al., 1982): Precambrian, Hercynian, and
Gondwana. The extremely long fetch (180˚–200˚ of long- Andean. Precambrian magmatism in the Huanuco
itude) available to circum-equatorial currents may have region consists of metaigneous ultramafic to mafic
contributed to higher latitude warming trends along the serpentinites, metagabbros, and metadiorites; syntectonic
Gondwana margin. metatonalites; and posttectonic dioritic and granitic
The global transgression, peaking during the Early intrusive bodies. Precambrian formations in the eastern
Permian, was probably not a relatively significant event Cordillera include two large outcrops in central Peru
(Vail et al., 1977). The transgressive highstand was the (6˚–13˚ S lat). The Huanuco region consists of granulites
lowest of the Paleozoic and indeed one of the lowest dated at about 600 Ma using U-Pb from zircons.
during the Phanerozoic. (For a discussion of exposed Magmatism within the Precambrian belt includes acid
continental land area during the Phanerozoic, see Tardy and basic synsedimentary rocks, ultrabasic and basic
et al., 1989.) The Early Permian can thus be considered as complexes, orthogneiss, and posttectonic intrusions.
the beginning of a long-term global warming trend in There is a lack of chemical analysis for the orthogneiss
Bolivia, which apparently persisted through the Eocene. bodies, but they are thought to be Precambrian–
Cambrian in age. The relationship between the post-
tectonic intrusions and the lower Paleozoic sequence is
unclear. Some plutons appear to be pre-Carboniferous,
IGNEOUS AND METAMORPHIC as they are nonconformably overlain by the Ambo
BASIN-BOUNDING FEATURES Group south of Chullay.
Precambrian events comprising the Brazilian shield of
Figure 12 summarizes Precambrian, Ordovician, eastern Bolivia (Figure 12) have been summarized by
Carboniferous, and Lower Permian igneous and meta- Litherland et al. (1985) and include gneisses, granulites
morphic rocks that occur in Peru, Bolivia, Argentina, and and other protoliths (>1400 Ma), the San Ignacio Schist
Chile. Diverse lithologies and events form the Precam- and plutons (about 1300 Ma), and the Rodonia–Sunsas
brian complexes that were apparently uplifted from time granites (1000 Ma). The most significant aspect of the
to time, thereby supplying sediment to the Bolivian and shield elements in Bolivia, however, is that they parallel
Peruvian basins. Brief descriptions of igneous and meta- the outboard Andean belts (Litherland et al., 1985). More
morphic rocks bounding the depositional settings are recent work by Salinas (1992) has shown evidence for
described here. some late Precambrian and Cambrian plutons and
Ordovician dikes and sills in Bolivia.
Arequipa Massif
Pampean Ranges of Argentina
According to Shackleton et al. (1979), there are discrete
igneous and depositional episodes within the Arequipa The Pampean cycle is described as being equivalent to
massif, as follows. Precambrian Mollendo metamor- the Brasiliano cycle farther east (Rapela et al., 1982). It
phism includes metasedimentary rocks, sillimanite- took place before 570 Ma, as evidenced by the age of
bearing gneiss, and staurolite-andalusite schists. The postkinematic plutons. The granitoids in the northern
244 Isaacson and Díaz Martínez

Figure 13—Cross-sectional model of the western Gondwana margin during the Early Devonian, including the San Nicolás
batholiths (Mukasa and Henry, 1990). It is suggested that crustal loading by the Arequipa (sialic) massif provided subsidence
for thick, clastic-dominated Devonian sedimentation, which further loaded the crust and continued subsidence.

part of the Pampean Ranges can be separated into three GEODYNAMIC MODEL
main types: pre- and synkinematic granitoids, late
kinematic granitoids, and postkinematic granitoids. The Figure 13 presents a preliminary southwest-northeast
Cafayate Granite (tonalite to granite) is the largest cross-sectional model of the central Andes during Early
granitoid unit in the Sierra de Quilmes and is interpreted Devonian time. Many aspects of the model require more
to have a 475-Ma emplacement age. The Cuchiyaco detailed information, including detailed biostratigraphy,
Granodiorite is a smaller pluton in the eastern part of the modeling of facies architecture, and consequent basin
Sierra de Quilmes. The Cerro Amarillo Granite is a post- analysis. The San Nicolás batholiths (Mukasa and Henry,
tectonic granitoid. K-Ar mineral ages indicate that most 1990) are a response to subduction and creation of a
of the units were intruded during the Paleozoic, magmatic arc, which comprised the Arequipa massif and
spanning 300 Ma. Low Sr ratios in various granitoid possibly other magmatic arcs to the south into Chile.
groups means that parental magma was derived from Attendant uplift of the Arequipa massif produced a
the upper mantle, accompanied by melting of local western land source that contributed much of the coarser
basement rocks. The main characteristic of the basement lithofacies in the adjacent foreland basin (central Andes).
in the Pampean Ranges is that no east-west control of The Paleozoic arc-trench gap is now missing, and its
granitoid emplacement is indicated (from Rb-Sr data). disposition has been debated extensively. It could have
There are four peaks of igneous activity detected: pre- been attritionally subducted during more active
and synkinematic bodies older than 500 Ma, late Mesozoic and Cenozoic subduction, or it could be part of
kinematic bodies (475 Ma), and two postkinematic units a rifted allochthonous terrane now comprising another
(440 and 340 Ma). region. Insufficient data are available to resolve this
problem.
Northern Chile Fragmentary evidence supporting the existence of a
foreland basin includes very thick, rapid siliciclastic sedi-
The presence of Precambrian rocks is not well docu- mentation along a La Paz “depocenter” (foredeep?)
mented in northern Chile. There are quartz-mica schists through late Paleozoic time, as well as periodic silling of
that are intruded by Ordovician (“Lila”) granites (450 subbasins (e.g., between the Belén and Ayo Ayo sections
Ma), thereby dating the protoliths as pre-Ordovician. The of Isaacson, 1977) by syndepositional basement block
Ordovician sedimentation and attendant volcanism faulting(?). Considering that Devonian siliciclastics were
ended with the Ocloyic deformation phase and its syn- mostly deposited in shallow water, the foreland may
kinematic granitic plutonism (Boric et al., 1990). There is have been a response to thrust fault loading on relatively
significant plutonism in northern Chile (318–225 Ma), thick crust. This would produce a large, shallow-water
marking Hercynian events there (Boric et al., 1990). foreland basin (Tankard, 1986).
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 245

Figure 14—Summary of stratigraphy, plutonism, biogeography, suggested relative sea levels, paleolatitude, and basin
dynamics for Bolivia basin during the Devonian–Early Permian.

An Early Devonian ephemeral intraarc basin with CONCLUSIONS


volcanogenic sediments and nonendemic faunas (Boucot
et al., 1980) appeared within the magmatic arc. Isolation Devonian, Carboniferous, and Permian stratigraphy,
from the foreland basin with its Malvinokaffric fauna lithofacies, and paleogeography are summarized in
was a result of the limits of the intraarc setting. Although Figure 14. The Devonian was influenced by a western
little basin analysis has been done on the Devonian of the land mass and significant subsidence of the Bolivia basin,
foreland basin, the following evidence supports its although details of facies architecture and sequence
existence. The deepening nature of the basin in Middle stratigraphy for the Devonian have yet to be developed.
Devonian time is suggested by rapid subsidence that The Devonian is dominated by siliciclastic units that
initially was tectonically induced. Possible crustal contain highly endemic Malvinokaffric fossils in its lower
loading and attendant transgression permitted accom- part and a mixed fauna in its middle part. Late Devonian
modation of additional siliciclastic sedimentation, some palynomorphs have been identified from the sequence,
of it sourced from the west. although how much of the Devonian is represented in
Basin activity during Late Devonian–middle Carbon- the stratigraphy has yet to be determined. Very few
iferous time is difficult to trace, although during Late megafossils of biostratigraphic utility occur in this part of
Carboniferous–Permian time, it appears that foreland the sequence. It appears evident, however, that faunas
subsidence continued. The Arequipa massif resumed its from North Africa and North America entered Bolivia
role as a source for coarse lithofacies in the Lake Titicaca from transgressions as the region progressively passed
area (Figure 8). A newer period of magmatic activity, through lower latitudes (Figure 14).
which eventually culminated in Late Permian plutonism Preliminary work has demonstrated the presence of a
and volcaniclastic deposition in much of the region, Late Devonian (Famennian) glaciomarine unit (diamic-
indicated resumption of subduction. This apparently tite) that indicates a proximal, locally glaciated (alpine?)
accelerated in the Mesozoic (James, 1971; Pitcher, 1984). region. Provenance analysis of clasts in the diamictite
246 Isaacson and Díaz Martínez

should identify source areas in the Cordillera Real western Argentina and northern Chile): Journal of South
and/or Arequipa massif. Early Carboniferous units American Earth Sciences, v. 4, p. 171–188.
reflect additional periglacial influences, along with Bahlburg, H., C. Breitkreuz, and W. Zeil, 1986, Palaozoische
fluvial and nearshore siliciclastics. An undated coal- Sedimente Nordchiles: Berliner geowissentschaftliche
Abhandlung (A), v. 66, 147–168.
bearing unit caps the Lower Carboniferous sequence,
Barrett, S. F., and P. E. Isaacson, 1988, Devonian Paleogeog-
indicating a lower latitude for the region. raphy of South America, in N. J. McMillan, A. F. Embry,
After a middle Carboniferous hiatus, during which and D. J. Glass, eds., Devonian of the World: Canadian
erosion removed a varying amount of Paleozoic units, Society of Petroleum Geologists Memoir 14, v. I, p. 655–667.
carbonates of the Late Carboniferous–Early Permian Barrios, L., and G. Beccar, 1988, Estratigrafía de detalle del área
indicate cyclic marine development, with subaerial Altiplano y Cordillera Oriental: Informe Interno,
features and western-sourced siliciclastics interrupting Yacimientos Petrolíferas Fiscales Bolivianos, Santa Cruz.
the shallow marine deposition. The faunas, as well as Barth, W., 1972, Das Permokarbon bei Zudañez (Bolivien) und
several primary features, demonstrate a low latitudinal eine Ubersicht des Jungpaläozoikums im zentralen Teil der
setting at about 20˚–25˚ S lat (Figure 14). Andes: Geologisches Rundschau, v. 61, p. 249–270.
Bell, C. M., 1987, The origin of the upper Palaeozoic Chañaral
Considerable emphasis has been placed on the
Melange of Northern Chile: Journal of the Geological
Arequipa massif as a source area for much of the Society of London, v. 144, p. 599–610.
Devonian System in Bolivia (especially western Bolivia), Bellido, E., and C. Guevara, 1963, Geologia de los cuadran-
and new evidence suggests that it was uplifted during at gulos de Punta de Bombon y Clemesi: Lima, Comision
least two times during the Paleozoic. It is also apparent Carta Geologia Nacional de Peru, Boletin no. 5, 92 p.
that two other times of significant orogenic activity Boric, R., F. Díaz, and V. Maksaev, 1990, Geologia y
occurred during the Paleozoic. Mechanisms for uplift Yacimientos Metaliferos de la Region de Antofagasta:
appear to have been the development of a magmatic arc Servicio Nacional de Geologia y Mineria de Chile, Boletin
during the Early Devonian, with possible crustal loading no. 40, 246 p.
through compression at this time. Boucot, A. J., 1971, Malvinokaffric Devonian marine
community distribution—implications for Gondwana:
Anais da Academia Brasileira de Ciencias, v. 43, p. 23–49.
Boucot, A. J., and J. Gray, 1979, Epilogue: a Paleozoic
Pangaea?, in J. Gray and A. J. Boucot, eds., Historical
Acknowledgments We acknowledge the donors of the
Biogeography, Plate Tectonics, and the Changing Environ-
Petroleum Research Fund (American Chemical Society), who ment: Corvallis, Oregon State University Press, p. 465–482.
provided support for fieldwork resulting in this paper. The Boucot, A. J., P. E. Isaacson, and G. Laubacher, 1980, An Early
National Geographic Society provided partial support for field Devonian, eastern Americas faunule from the coast of
data presented here. Through assistance from the Servicio southern Peru: Journal of Paleontology, v. 54, p. 359–365.
Geológico de Bolivia (GEOBOL) and Yacimientos Petrolíferos Breitkreuz, C., and H. Bahlburg, 1985, Paleozoic flysch series in
Fiscales Bolivianos (YPFB), including extensive discussions the coastal Cordillera of northern Chile: Geologische
with J. C. Lema, R. Suárez S., A. Dalenz, and others, we have Rundschau, v. 74, p. 565–572.
been able to continue fieldwork in Bolivia. J. Fay assisted with Brodzikowsky, K., and A. J. Van Loon, 1991, Glacigenic
the compilations of igneous rock information in the region. sediments: Developments in Sedimentology, v. 49,
Amsterdam, Elsevier, 674 p.
Caputo, M. V., 1985, Late Devonian glaciation in South
America: Palaeogeography, Palaeoclimatology, Palaeoe-
REFERENCES CITED cology, v. 51, p. 291–317.
Caputo, M. V., and J. C. Crowell, 1985, Migration of glacial
centers across Gondwana during Paleozoic era: GSA
Ahfleld, F., and L. Branisa, 1960, Geologia de Bolivia: Editorial
Bulletin, v. 96, p. 1020–1036.
Don Bosco, La Paz, 245 p.
Carlier, G., G. Grandin, G. Laubacher, R. Marocco, and F.
Allmendinger, R. W., V. A. Ramos, T. E. Jordan, M. Palma, and
Mégard, 1982, Present knowledge of the magmatic
B. L. Isacks, 1983, Paleogeography and Andean structural
evolution of the eastern Cordillera of Peru: Earth Science
geometry, northwest Argentina: Tectonics, v. 2, p. 1–16.
Reviews, v. 18, p. 253–283.
Azcuy, C. L., and R. Suárez-Soruco, 1984, El género Nothorha-
Clarke, J. M., 1913, Fosseis Devonianos do Paraná: Monographs,
copteris en el Paleozoico superior de la Península de
Serviço Geologico e Mineralogico do Brasil, v. I, 353 p.
Copacabana, Bolivia: Bulletin, Project IGCP 211, p. 45–46.
Cobbing, E. J., J. M. Ozard, and N. J. Snelling, 1977, Reconnais-
Azcuy, C. L., G. Laffite, and L. A. Rodrigo, 1982, El limite
sance geochronology of the crystalline basement rocks of
Carbonico-Permico en la Cuenca Tarija-Titicaca: Actas, III
the coastal Cordillera of Peru: GSA Bulletin, v. 88,
Congreso Argentino de Paleontologia y Bioestratigrafia,
p. 241–246.
Corrientes, p. 39–44.
Coira, B., J. Davidson, C. Mpodozis, and V. Ramos, 1982,
Babin, C., P. R. Racheboeuf, A. Le Hérissé, and M. Suárez
Tectonic and magmatic evolution of the Andes of northern
Riglos, 1991, Donées Nouvelles sur les Goniatites du
Argentina and Chile: Earth Science Reviews, v. 18,
Dévonien de Bolivia: Geobios, v. 24, p. 719–724.
p. 303–332.
Bahlburg, H., 1985, Sedimentological aspects of the El Toco
Crowell, J. C., 1978, Gondwanan glaciation, cyclothems, conti-
Formation (Paleozoic; Coastal Cordillera) NW of Quillagua,
nental positioning, and climatic change: American Journal
northern Chile: Actas, Congreso Geologico Chileno No. 4,
of Science, v. 278, p. 1345–1372.
V. 1, p. 17–28.
Dalmayrac, B., 1977, Geologie des Andes Peruviennes,
Bahlburg, H., and C. Breitkreuz, 1991, Paleozoic evolution of
Geologie de la Cordillere Orientale de la Region de
active margin basins in the southern Central Andes (north-
Huanuco; sa Place dans une Transversale des Andes du
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 247

Perou Central (9˚S, 10˚30’S): Ph.D. dissertation, Université Erwin, D. M., H. W. Pfefferkorn, V. Alleman, and H. Nuñez
des Sciences et Techniques du Languedoc, Montpellier, del Prado, 1992, Warm temperate climatic Realm on the
123 p. paleo-Pacific margin of Late Carboniferous Gondwanaland
Dalmayrac, B., G. Laubacher, R. Marocco, C. Martínez, and P. (Peru): GSA Annual Meeting, Abstracts with Programs, v.
Tomasi, 1980, La chaine hercynienne d’Amerique du sud: 24, p. A192.
estructure et evolution d’un orogene intracratonique: Geol- Esteban, M., and C. F. Klappa, 1983, Subaerial exposure envi-
ogische Rundschau, v. 69, p. 1–21. ronment, in P. A. Scholle, D. G. Bebout, and C. H. Moore,
da Silva, O. B., 1988, Revisao Estratigrafica da Bacia do eds., Carbonate depositional environments: AAPG Memoir
Solimoes: Anais, XXXV Congresso Brasileiro de Geologia, 33, p. 1–54.
Belém, Pará, v. 6., p. 2428–2438. Flügel, E., 1982, Microfacies Analysis of Limestones: Berlin,
Davidson, J., C. Mpodozis, and S. Rivano, 1981, El Paleozoico Springer-Verlag, 633 p.
de Sierra de Almeida, al oeste de Monturaqui, alta Frutos, J., and A. Tobar, 1975, Evolution of the southwestern
cordillera de Antofagasta, Chile: Revista Geologica de continental margin of South America, in K. S. W. Campbell,
Chile, no. 12, p. 3–23. ed., Gondwana geology: Third Gondwana Symposium,
deMelo, J. H. G., 1988, The Malvinokaffric Realm in the Australian National University Press, p. 565–578.
Devonian of Brazil, in N. J. McMillan, A. F. Embry, and D. J. Gobbett, D. J., 1973, Permian Fusulinacea, in A. Hallam, ed.,
Glass, eds., Devonian of the World: Canadian Society of Atlas of Palaeobiogeography: Amsterdam, Elsevier,
Petroleum Geologists Memoir 14, v. I, p. 669–703. p. 151–158.
Díaz, E., 1991, Litoestratigrafía del Carbonífero del Altiplano Hambrey, M., and W. Harland, 1981. Earth’s Pre-Pleistocene
de Bolivia: Revista Técnica de Yacimientos Petrolíferas Glacial Record: Cambridge, Cambridge University Press,
Fiscales Bolivianos, v. 12, p. 295–302. 1044 p.
Díaz, E., and J. C. Lema, 1991a, Diamictitas glaciomarinas en el Havlicek, V., and P. Röhlich, 1987, Devonian and Carbonif-
Carbonífero del Altiplano norte de Bolivia: Sedimentología erous brachiopods from the northern flank of the Murzuq
e interpretación de ambientes sedimentarios: Actas, 6º basin (Libya): Sborník geologicky ved, Paleontology, v. 28,
Congreso Geológico de Chile, Viña del Mar, v. I, p. 268–271. p. 117–177.
Díaz, E., and J. C. Lema, 1991b, Revisión litoestratigráfica del Hervé, F., J. Davidson, E. Godoy, C. Mpodozis, and V. Covace-
Carbonífero del Altiplano norte de Bolivia: Actas, 6º vich, 1981, The late Paleozoic in Chile: stratigraphy,
Congreso Geológico de Chile, Viña del Mar, v. I, p. 574–578. structure, and possible tectonic framework: Anais,
Díaz, E., P. E. Isaacson, K. L. Christensen, M. Vavrdová, and B. Academia brasiliensis Ciencias, v. 53, p. 361–373.
Antelo Pérez, 1993, Significance of Late Devonian diamic- Hill, D., 1981, Rugosa and Tabulata, in Treatise on Invertebrate
tite, Lake Titikaka, Bolivia: Comptes Rendus, XII Interna- Paleontology, Part F, Coelenterata: GSA and University of
tional Congress of Carboniferous and Permian Stratigraphy Kansas Press, Lawrence, supplement 1, 2 vols, 762 p.
and Geology, Buenos Aires, v. 1, p. 293–304. Hünicken, M. A., J. H. G. deMelo, and V. B. Lemos, 1988,
Díaz, H., 1959, Comunicación acerca de las condiciones Devonian conodonts from the Upper Amazon basin, north-
geológicas en el curso superior del; río Beni: Boletín Técnico western Brazil, in N. J. McMillan, A. F. Embry, and D. J.
de Yacimientos Petroliferos Fiscales Bolivianos, v. 1, Glass, eds., Devonian of the world: Canadian Society of
p. 21–32. Petroleum Geologists Memoir 14, v. III, p. 479–483.
Díaz Martinez, E., P. E. Isaacson, and P. E. Sablock, 1992, Late Isaacson, P. E., 1974, First South American occurrence of
Paleozoic latitudinal shift of Gondwana: stratigraphic, sedi- Globithyris: its ecological and age significance in the Malvi-
mentologic and biogeographic evidence from Bolivia: Table nokaffric Realm: Journal of Paleontology, v. 48, p. 778–784.
Ronde Européen Paléontologie et Stratigraphie d’Amérique Isaacson, P. E., 1975, Evidence for a western extracontinental
Latine, Documents, Laboratoire Géologie, Lyon, p. 119–138. land source during the Devonian Period in the central
Dickinson, W. R., 1988. Provenance and sediment dispersal in Andes: GSA Bulletin, v. 86, p. 39–46.
relation to paleotectonics and paleogeography of sedimen- Isaacson, P. E., 1977, Devonian stratigraphy and brachiopod
tary basins, in K. L. Kleinspehn and C. Paola, eds., New paleontology of Bolivia, part A: Orthida and Strophome-
Perspectives in Basin Analysis: New York, Springer-Verlag, nida: Palaeontographica, Abteilung A, v. 155, p. 133–192.
p. 3–25. Isaacson, P. E., 1989, Late(?) Devonian orogeny in Peru: corre-
d’Orbigny, A., 1842, Voyages dans l’Amerique meridionale de lation to Antler events in western North American
1826–1833: Paleontologie, v. 3, part 4, 188 p. cordillera: GSA Meeting, Cordilleran and Rocky Mountain
Dowdeswell, J. A., 1988, A model for iceberg sedimentation in Sections, Abstracts with Programs, v. 21, p. A97.
varying glaciological and oceanographic settings: GSA Isaacson, P. E., and D. G. Perry, 1977, Biogeography and
Annual Meeting, Abstracts with Programs, v. 20, p. A84. morphological conservatism of Tropidoleptus (Brachiopoda,
Duarte, R. G., 1989, Estratigrafía de Detalle del Formacion Orthida) during the Devonian: Journal of Paleontology, v.
Copacabana y Mesozoico Inicial: Formaciones Sayari y 51, p. 1108–1122.
Ravelo desl Area Oeste de Cochabamba: Tésis de Grado, Isaacson, P. E., and P. E. Sablock, 1988, Devonian System in
Universidada Mayor des San Andrés, La Paz, Bolivia, 89 p. Bolivia, Peru, and northern Chile, in N. J. McMillan, A. F.
Dunbar, C. O., and N. D. Newell, 1946, Marine Early Permian Embry, and D. J. Glass, eds., Devonian of the world:
of the central Andes and its fusuline faunas: American Canadian Society of Petroleum Geologists Memoir 14, v. I,
Journal of Science, v. 244, p. 377–402. p. 719–728.
Dutro, J. T., Jr., and P. E. Isaacson, 1991, Paleogeographic Isaacson, P. E., and P. E. Sablock, 1990, Devonian palaeogeog-
significance of circum-Gondwana Lower Carboniferous raphy and palaeobiogeography of the central Andes, in W.
“big shell” brachiopod associations (abs.): XII International S. McKerrow and C. R. Scotese, eds., Palaeozoic palaeo-
Congress on the Carboniferous and Permian, Buenos Aires, geography and biogeography: Geological Society of
p. 33–34. London Memoir 12, p. 431–435.
Eckel, E. B., 1959, Geology and mineral resources of Isaacson, P. E., J. Davidson, and L. L. Fisher, 1985, Devonian
paraguay–a reconnaissance: U.S.G.S. Professional Paper and Carboniferous stratigraphy of the Sierra de Almeida,
327, 110 p. northern Chile, preliminary results: Revista Geologica de
248 Isaacson and Díaz Martínez

Chile, no. 25-26, p. 113–121. Université des Sciences et Techniques du Languedoc,


Isaacson, P. E., K. L. Canter, and P. E. Sablock,1993, Late Montpellier, 77 p.
Paleozoic Copacabana Formation in N.W. Bolivia: Paleo- Miller, H., 1970, Das Problem des hypothetischen “Pazifischen
geographic significance of carbonates with siliciclastics: Kontinentes,” gesehen von der chilenischen Pazifikkuste:
Comptes Rendus, XII International Congress on the Geologische Rundschau, v. 59, p. 927–938.
Carboniferous and Permian, Buenos Aires, v. 2, p. 261–268. Molnia, B. F., 1983, Glacial-Marine Sedimentation: New York,
Isaacson, P. E., B. P. Palmer, B. L. Mamet, J. C. Cooke, and D. E. Plenum Press, 844 p.
Sanders, 1995, Devonian–Carboniferous stratigraphy in the Molnia, B. F., 1988, Comparison of glacial-marine depositional
Madre de Dios basin, Bolivia: Pando X-1 and Manuripi X-1 environments of polar Antarctica and the temperate Gulf of
wells, in A. J. Tankard, R. Suarez, and H. J. Welsink, Alaska: GSA Meeting, Abstracts with Programs, v. 20,
Petroleum basins of South America: AAPG Memoir 62, this p. A84.
volume. Mosmann, R., F. U. Falkenheim, A. Gonçalves, and F. N. Filho,
James, D. E., 1971, Plate tectonic model for the evolution of the 1986, Oil and gas potential of the Amazon Paleozoic basins,
central Andes: GSA Bulletin, v. 82, p. 3325–3346. in M. T. Halbouty, ed., Future petroleum provinces of the
Johnson, D. A., 1974, Deep Pacific circulation: intensification world: AAPG Memoir 40, p. 207–241.
during the early Cenozoic: Marine Geology, v. 17, p. 71–78. Mpodozis, C., and R. Forsythe, 1983, Stratigraphy and
Laubacher, G., 1974, Le Paleozoique Inferieur de la Cordillere geochemistry of accreted fragments of the ancestral Pacific
Orientale du SE du Perou: Cahiers Office de la Recherche floor in southern South America: Palaeogeography, Palaeo-
Scientifique et Technique Outre-Mer, serie Geologie, Paris, climatology, Palaeoecology, v. 41, p. 103–124.
v. VI, p. 29–40. Mukasa, S. B., and D. J. Henry, 1990, The San Nicolás batholith
Laubacher, G., 1977, Geologie des Andes Peruviennes: of coastal Peru: early Palaeozoic continental arc or conti-
Geologie de l’Altiplano et de la Cordillere Orientale au nental rift magmatism?: Journal of Geological Society of
nord et nord-ouest du Lac Titicaca (Perou): Ph. D. disserta- London, v. 147, p. 27–39.
tion, Université des Sciences et Techniques du Languedoc, Newell, N. D., J. Chronic, and T. G. Roberts, 1953, Upper
Montpellier, 117 p. Paleozoic of Peru: GSA Memoir 58, 272 p.
Laubacher, G., A. J. Boucot, and J. Gray, 1982, Additions to Niemeyer, H., F. Urzua, F. G. Aceñolaza, and C. R. González,
Silurian stratigraphy, lithofacies, biogeography and paleon- 1985, Progresos recientes en el conocimiento del Paleozoico
tology of Bolivia and southern Peru: Journal of Paleon- de la region de Antofagasta: IV Congreso Geologico
tology, v. 56, p. 1138–1170. Chileno, v. 1, p. 410–435.
Le Hérissé, A., F. Paris, and P. R. Racheboeuf, 1992, Marine Palma, M. A., P. O. Parica, and V. A. Ramos, V. A., 1987, El
palynomorphs from the Devonian of Bolivia, in Le Dévoni- Granitico Archibarca: su edad y significado tectonico,
enne de Bolivie, Le Cadre Stratigraphique Revu a la Provincia de Catamarca: Revista, Asociacion Geologica
Corrélations Intercontinentales: Résumés, Paléontologie et Argentina, v. 41, p. 414–419.
Stratigraphie d’Amerique Latine, Table Ronde Européene, Pitcher, W. S., 1984, Phanerozoic plutonism in the Peruvian
Lyon, p. 32. Andes, in R. S. Harmon and B. A. Barreiro, eds., Andean
Litherland, M., B. A. Klinck, E. A. O’Connor, and P. E. J. magmatism: Nantwich, England, Shiva Publishing,
Pitfield, 1985, Andean-trending mobile belts in the Brazilian p. 152–167.
shield: Nature, v. 314, p. 345–348. Powell, R. D., 1988, Sediment accumulation rates as second
Lobo-Boneta, J., 1987, El Límite Devónico–Carbonifero en el order controls of glacial fluctuations of tidewater fronts:
Subandino Sur de Bolivia: Revista Técnica de Yacimientos GSA Annual Meeting, Abstracts with Programs, v. 20, no.
Petroliferos Fiscales Bolivianos, v. 10, no. 3-4, p. 213–217. 7, p. A85.
Mack, G. H., 1984, Exceptions to the relationship between plate Racheboeuf, P. R., A. Le Hérissé, C. Babin, F. Guillocheau, and
tectonics and sandstone composition: Journal of Sedimen- M. Truyols Massoni, 1992, Le Dévonienne de Bolivie, in Le
tary Petrology, v. 54, no. 1, p. 212–220. Cadre Stratigraphique Revu a la Corrélations Intercontinen-
Marocco, R., 1977, Geologie des Andes Peruviennes, un tales: Résumés, Paléontologie et Stratigraphie d’Amerique
Segment E.W. de la Chaine des Andes Peruviennes: la Latine, Table Ronde Européene, Lyon, p. 42.
Deflexion d’Abancay—Etude Geologique de la Cordillere Ramos, V. A., T. E. Jordan, R. W. Allmendinger, C. Mpodozis,
Orientale et des Hauts Plateaux entre Cuzco et San Miguel S. M. Kay, J. M. Cortes, and M. A. Palma, 1986, Paleozoic
(Sud du Perou 12˚30’S a 14˚00S): Ph.D. dissertation, Univer- terranes of the central Argentine-Chilean Andes: Tectonics,
sité des Sciences et Techniques du Languedoc, Montpellier, v. 5, p. 855–880.
141 p. Rapalini, A., and J. Vilas, 1991, Paleomagnetismo, in El Sistema
Marocco, T. Sempere, D. Merino, J. Oller, M. Pérez, and E. Permico en la Republica Argentina y en la Republica
Soria, 1987, Le Permo-Carbonifère du Lac Titicaca (nord de Oriental del Uruguay: XII Congreso Internacional de la
la Bolivie): un exemple d’inversion de polarité dans un Estratigrafía y Geología del Carbonifero y Pérmico y
bassin: Séminaire “Géodinamique des Andes Centrales,” Academia Nacional de Ciencias de Córdoba, p. 18–25.
Résumés des Communications, Orstom, Paris, p. 48–51. Rapela, C. W., L. M. Heaman, and R. H. McNutt, 1982, Rb-Sr
Martínez, C., 1980, Structure et évolution de la chaine hercyni- geochronology of granitoid rocks from the Pampean
enne et de la chaine andine dans le nord de la Cordillère Ranges, Argentina: Journal of Geology, v. 90, p. 574–582.
des Andes de Bolivie: Travaux Doctor, Orstom, Paris, v. Raymond, A., 1990, Dead by degrees: articulate brachipods,
119, 352 p. paleoclimate, and the middle Carboniferous extinction
Matsch, C. L., and R. W. Ojakangas, 1988, Comparisons in event: Palaios, v. 5, p. 111–123.
depositional style of polar and temperate glacial ice: late Raymond, A., P. H. Kelley, and C. B. Lutken, 1989, Polar
Paleozoic Whiteout Conglomerate (West Antarctica) and glaciers and life at the equator: the history of Dinantian and
late Proterozoic Mineral Forks Tillite (Utah): GSA Annual Namurian (Carboniferous) climate: Geology, v. 17, p.
Meeting, Abstracts with Programs, v. 20, p. A133. 408–411.
Mégard, F., 1973, Etude geologique d’une transversale des Richter, R., and E. Richter, 1942, Die Trilobiten der Weismes-
Andes au niveau du Perou central: Ph.D. dissertation, Schichten am Hohen Venn, mit Bemerkungen uber die
Evidence for a Middle–Late Paleozoic Foreland Basin, Paleolatitudinal Shift, Central Andes 249

Malvinocaffrische Provinz: Senckenbergiana, v. 25, Starck, D., E. Gallardo, and A. Schulz, 1993, Neopaleozoic
p. 156–179. stratigraphy of the Sierras Subandinas occidentales and
Rivano, S., and P. Sepúlveda, 1983, Hallazgo de Foraminíferos Cordillera Oriental Argentina, with comments on the
del Carbónifero Superior en la Formación Huentelauquen: southern borders of the Tarija basin: XII International
Revista Geologica de Chile, no. 19–20, p. 25–35. Congress on the Carboniferous and Permian, Buenos Aires,
Rocha Campos, A. C., 1983, North Andean Area, in C. v. 2, p. 353–372.
Martinez Díaz, ed., The Carboniferous of the world: Suárez-Riglos, M., M. A. Hünicken, and D. Merino, 1987,
Instituto Geologico y Minero de España, v. II, p. 180–200. Conodont biostratigraphy of the Upper
Rodrigo, L. A., and A. Castaños, 1978, Sinopsis estratigráfica Carboniferous–Lower Permian rocks of Bolivia, in R. L.
de Bolivia (1): Paleozoico: Academia Nacional de las Austin, ed., Conodonts: investigative techniques and appli-
Ciencias de Bolivia, La Paz, 146 p. cations: British Micropalaeontological Society, Chichester,
Roeder, D., and R. Chamberlain, 1995, Structural geology of Ellis Horwood Publishers, p. 317–325.
sub-Andean fold and thrust belt in northwestern Boliva, in Suárez-Soruco, R., and J. Lobo-Boneta, 1983, La Fase Compre-
A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum siva Eohercinica en el Sector Oriental de la Cuenca
basins of South America: AAPG Memoir 62, this volume. Cordillerana de Bolivia: Revista Técnica de Yacimientos
Rösler, O., R. Iannuzzi, and R. Súarez-Soruco, 1989, A Flora Petrolíferas Fiscales Bolivianos, v. 9, p. 189–202.
Carbonifera da Formaçao Kasa em Belem, Península de Suttner, L. J., A. Basu, A. and G. H. Mack, 1981, Climate and
Copacabana, Altiplano Boliviano, e a Importancia das the origin of quartz arenites: Journal of Sedimentary
Formas Trifoliadas: XI Congresso Brasileiro de Paleon- Petrology, v. 51, no. 4, p. 1235–1246.
tologia, Resumo das Comunicaçoes, p. 45–46. Tankard, A. J., 1986, On the depositional response to thrusting
Ross, C. A., 1973, Carboniferous Formaninferida, in A. Hallam, and lithospheric flexure: examples from the Appalachian
ed., Atlas of Palaeobiogeography: Amsterdam, Elsevier, and Rocky Mountain basins: Special Publications, Interna-
p. 127–132. tional Association of Sedimentologists, v. 8, p. 369–392.
Russo, A., 1966, Algunas consideraciones fisiográficas del terri- Tardy, E., R. N’Kounkou, and J. Probst, 1989, The global water
torio boliviano: Boletín, Instituto Boliviano del Petróleo, cycle and continental erosion, during Phanerozoic time (570
v. 6, p. 7–25. m.y.): American Journal of Science, v. 289, p. 455–483.
Sablock, P. E., 1993, A warm to cold paleoclimate change in Vail, P. R., R. M. Mitchum, and S. Thompson, 1977, Global
Gondwana (Devonian through Carboniferous): a conse- cycles of relative changes of sea level, in C. E. Payton, ed.,
quence of coupled continental shift and onlap: Comptes Seismic stratigraphy—applications to hydrocarbon explo-
Rendus, XII International Congress on the Carboniferous ration: AAPG Memoir 26, 515 p.
and Permian, Buenos Aires, v. 1, p. 355–367. Vavrdová, M., P. E. Isaacson, E. Díaz Martinez, and J. Bek,
Sakagami, S., J. Yanagida, T. Ishibashi, T. Kawabe, T. Kase, K. 1991, Palinologia del Limite Devonico-Carbonifero en torno
Nagai, T. Sugiyama, R. Carrasco, A. Escobar, and C. al Lago Titikaka, Bolivia: Resultados Preliminares: Revista
Rangel, 1986, Biostratigraphic study of Paleozoic and Tecnica de Yacimientos Petrolíferas Fiscales Bolivianos, v.
Mesozoic groups in Central Andes—an interim report (2): 12, p. 303–313.
Department of Earth Sciences, Faculty of Science, Chiba Vavrdová, M., P. E. Isaacson, E. Díaz Martinez, and J. Bek, in
University, Chiba, Japan, 83 p. press, Devonian–Carboniferous boundary at Lake Titikaka,
Salinas, M. A., 1992, El magmatismo Cámbrico–Ordovícico en Bolivia: preliminary palynological results: Comptes
Bolivia, in J. G. Gutiérrez Marco, J. Saavedra, and I. Rábano, Rendus, XII International Congress on the Carboniferous
eds., Paleozoico Inferior de Ibero-América: Universidad de and Permian, Buenos Aires, v. 1.
Extramadura, Salamanca, Spain, p. 241–253. Veevers, J. J., 1984, Phanerozoic Earth History of Australia:
Saunders, W. B., and W. H. C. Ramsbottom, 1986, The mid- Oxford, U.K., Clarendon Press, 418 p.
Carboniferous eustatic event: Geology, v. 14, p. 208–212. Veevers, J. J., and C. McA. Powell, 1987, Late Paleozoic glacial
Sempere, T., 1987, Caracteres geodinámicos generales del Pale- episodes in Gondwanaland reflected in
ozoico superior de Bolivia, in Late Paleozoic of South transgressive–regressive depositional sequences in
America: Fourth Annual Meeting, Santa Cruz, IGCP no. Euroamerica: GSA Bulletin, v. 98, p. 475–487.
211, p. 9–19. Wilson, E. C., 1990, Permian corals of Bolivia: Journal of Pale-
Sempere, T., 1989, Paleozoic evolution of the Central Andes ontology, v. 64, p. 60–78.
(10–26˚): 28th International Geological Congress, Abstracts, Zeil, W., 1979, The Andes—a geological review: Berlin,
Washington D.C., v. 3, p. 73. Gebrüder Borntraeger, 260 p.
Sempere, T., 1990, Cuadros estratigráficos de Bolivia: prop- Ziegler, P. A., 1989, Evolution of Laurussia: Boston, Kluwer
uestas nuevas: ORSTOM, v. 20, 26 p. Academic Publishers, 105 p.
Sempere, T., R. Marocco, D. Merino, J. Oller, M. Pérez, and E.
Soria, 1986, Los caracteres geodinámicos generales de la
sedimentación permo-carbónica al sur del Lago Titicaca: 8˚
Congreso Geológico de Bolivia, La Paz, p. 44.
Sempere, T., G. Herail, and J. Oller, 1988, Los aspectos estruc-
turales y sedimentarios del oróclino boliviano: 5˚ Congreso
Geologico Chile, Santiago, v. 1, p. A127–A142. Authors’ Mailing Address
Shackleton, R. M., A. C. Ries, M. P. Coward, and P. R.
Cobbold, 1979, Structure, metamorphism, and P. E. Isaacson
geochronology of the Arequipa masssif of coastal Peru: E. Díaz Martínez
Journal of Geological Society of London, v. 136, p. 195–214. Department of Geology
Sheffels, B. M., 1990, Lower bound on the amount of crustal University of Idaho
shortening in the central Bolivian Andes: Geology, v. 18, Moscow, Idaho 83843
p. 812–825. U.S.A.
Tectonic Evolution of the Andes of Northern Argentina
R. Mon J. A. Salfity
Universidad Nacional de Tucumán–CONICET Universidad Nacional de Salta–CONICET,
Tucumán, Argentina Salta, Argentina

Abstract

S everal superimposed tectonic stages distinguished by varying structural styles are recognized in the
Andes of northern Argentina (22˚–28˚ S lat). The oldest structures occur in the Precambrian crystalline
basement. This basement forms the central core of the region and is made up of several multiply deformed
belts. These belts were intruded by several generations of granitoids and were amalgamated during the
Panamerican orogeny (Late Brazilian orogeny, 700–600 Ma). A westward-vergent foldbelt containing Ordovi-
cian marine sediments shows eastward-dipping axial plane cleavage. It lies along the western border of the
crystalline core and acts as host rock to pretectonic intrusives. Development of the folding is assigned to the
Late Ordovician Ocloyic orogeny. The sub-Andean ranges and Puna Silurian–Devonian successions were
folded during the Late Devonian–Early Carboniferous at the beginning of the Gondwanan cycle (Chañic
orogeny). This tectonic cycle is represented in several late Paleozoic basins that surround the study area. The
inversion of those basins probably took place during the middle Permian San Rafael orogeny.
The Andean cycle commenced with the opening of rift troughs filled with thick continental deposits during
Early Cretaceous–Eocene time. The inversion of these troughs began with the late Eocene Inca movements, but
was completed during the Miocene Quechua and Pliocene–Pleistocene Diaguita orogenies. From late
Oligocene time onward, continental basins developed, and an extensive Miocene–Pleistocene volcanic arc orig-
inated on the western flank of the study area. These Cretaceous–Cenozoic basins were inverted by the Diaguita
orogeny. Andean tectonics caused clearly differentiated morphostructural units. The most westerly of these is
the Puna Plateau, characterized by Precambrian basement and Paleozoic rocks sheets that were thrust over
Tertiary continental successions. East of the Puna, the Eastern Cordillera represents a tectonic stack of Precam-
brian basement and Paleozoic rock sheets thrust eastward over the sub-Andean ranges. This latter belt forms
the outermost unit, made up of large faulted anticlines. South of 27˚ S lat, a change occurs in the architecture of
the Andean foreland. The sub-Andean ranges and the Eastern Cordillera are replaced by faulted blocks of
Precambrian crystalline basement and Paleozoic granitic intrusions, which form the Pampean ranges. This
paper summarizes the evolution of the oil-bearing basins of northern Argentina.

Resumen

E n los Andes del noroeste argentino (22˚–28˚ S) se distinguen varios pisos superpuestos con distintos
estilos estructurales. Las estructuras más antiguas se advierten en el basamento cristalino precámbrico,
núcleo central de la región, constituido por varios cinturones polideformados. Estos fueron intruidos por varias
generaciones de granitoides y amalgamados durante la orogenia Panamericana (700–600 Ma) (Brasiliano
superior). Un cinturón plegado constituido por sedimentos ordovícicos marinos, y por intrusivos pretec-
tónicos, que muestra clivaje inclinado hacia el este, se encuentra a lo largo del borde occidental del núcleo
cristalino. El desarrollo de este plegamiento se atribuye a la orogenia Oclóyica del Ordovícico superior. Los
depósitos del Silúrico-Devónico de la Puna y de las Sierras Subandinas fueron plegados durante el Devónico
superior-Carbonífero inferior al principio del ciclo Gondwánico (orogenia Cháñica). Este ciclo tectónico está
representado en varias cuencas del Paleozoico superior que rodean el área estudiada. La inversión de estas
cuencas se produjo por efecto de la orogenia San Rafael durante el Pérmico medio.
El ciclo Andino comenzó con la apertura de cuencas elongadas rellenadas con depósitos continentales del
Cretácico-Eoceno inferior. La inversión de estas cuencas comenzó con la orogenia Incaica en el Eoceno tardío,
pero se completó con las orogenias Quechua del Mioceno y Diaguita del Plioceno-Pleistoceno. A partir del
Oligoceno tardío se desarrollaron las cuencas neógenas y desde el Mioceno al Pleistoceno evolucionó un arco
volcánico en el flanco occidental del área estudiada. Estas cuencas cenozoicas fueron invertidas por la orogenia
Diaguita. La tectónica Andina originó unidades morfoestructurales claramente diferenciadas. La más occi-
dental es la Puna, caracterizada por láminas de basamento precámbrico y de Paleozoico corridas sobre las
rocas continentales terciarias. Hacia el este de la Puna la Cordillera Oriental representa un apilamiento
tectónico de escamas de basamento precámbrico y de rocas paleozoicas corridas sobre el cinturón de las Sierras
Subandinas. Este último forma la unidad más externa, constituida por grandes anticlinales fallados. Al sur de
27˚ S ocurre un cambio en la arquitectura del antepaís andino. Las Sierras Subandinas y la Cordillera Oriental
son reemplazadas por bloques fallados de basamento cristalino y de plutones graníticos, que constituyen las
Sierras Pampeanas. Este artículo incluye una reseña de la evolución de las cuencas productoras de hidrocar-
buros del norte argentino.

Mon, R., and J. A. Salfity, 1995, Tectonic evolution of the Andes of northern Argentina, in 269
A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America:
AAPG Memoir 62, p. 269–283.
270 Mon and Salfity

Figure 2—Summary chart of Phanerozoic geologic history


of northern Argentina. Column locations are shown in
Figure 1. Key: 1, continental facies; 2, deep marine facies;
3, marine shelf facies; 4, stable shelf facies; 5, diamictite
facies; 6, volcanics; 7, stratovolcano; 8, granitoid; 9, meta-
morphics; 10a, source rock interval; 10b, potential source
rock interval; 11, hydrocarbon reservoir; 12, coarsening-
upward sequence; 13, fining-upward sequence; 14, base or
top not seen; 15, start of sedimentary cycle; 16, unconfor-
mities (a > b > c); 17, diastrophic events.

a highly complex basement that forms the roots of the


present-day Andes. From Cambrian to Tertiary time, the
Figure 1—Location map of the study area showing the Phanerozoic basins underwent varying degrees of defor-
geologic provinces. The numbers indicate the location of mation and shortening.
columns shown in Figure 2. Key: a, Precambrian; b,
The aim of this study is to describe the evolution of
Neogene volcanic arc. Heavy line shows the location of
Figure 3. the Phanerozoic basins and the tectonic styles produced
by the orogenies affecting this region (Figure 2). This
study emphasizes the importance of the late Precam-
INTRODUCTION brian–early Paleozoic structural framework in guiding
the subsequent development of basins throughout the
Northwestern Argentina straddles the eastern Phanerozoic. Tankard et al. (1995) have expanded on this
Andean slope (Figure 1). It has been under intense theme on a regional scale. However, the manner in
tectonic activity from Precambrian time until the present which these older fabrics were reactivated depended on
day. This complex tectonism generated sedimentary their orientation relative to successive stress fields.
basins of different ages, both in the foreland and in the The evolution of the basins of northern Argentina is
interior of the Andean chain. The respective sedimentary intracratonic, and the Precambrian crystalline basement
successions are more than 15 km thick. The superposi- was intensely involved in the Andean structures. The
tion of diastrophism occurred at different crustal levels in Mesozoic and Cenozoic magmatic activity was not as
Tectonic Evolution of the Andes of Northern Argentina 271

intense as along the western slope of the Andes in The Precambrian crystalline basement consists of
northern Chile. Only the western border of the region several belts having different lithologic and structural
has been affected by a Neogene volcanic arc. characteristics (Figure 4). Low-grade schists predominate
The basins in northern Argentina that are most along its eastern border where three belts of mutually
prospective for oil exploration are situated north of 26˚ S discordant structural styles are observed. All of them are
lat. This area was studied in most detail, although the covered unconformably by Cambrian and Ordovician
maps and regional considerations were extended to 28˚ S beds. The Puncoviscana-Lules belt is a narrow external
lat (Figures 1, 3) to give a broader overview. Hydro- strip, in places only slightly deformed, and with a
carbon exploration is taking place in the productive predominantly westward vergence. The Choromoro belt
Tarija, Olmedo, and Metán basins. Thus far, exploration is located farther west and shows a greater degree of
in the Tres Cruces, Alhuampa, and Paganzo basins has deformation, eastward vergence, and kink folds with
not been successful. partially boudinated limbs. Its deformation is older than
600 Ma, and it is intruded by postkinematic granitoids
with isotopic ages near 600 Ma (Halpern and Latorre,
PRECAMBRIAN BASEMENT 1973; Omarini et al., 1985). Between 26˚ and 27˚ S lat, the
Aguas Blancas belt appears as a stretch of westward
The Precambrian basement lies between 23˚ and 34˚ S vergent and intensely microfolded schists.
lat (Figures 1, 3, 4) and separates the Andean chain from The central part of the crystalline basement core
the Chaco plain and the Chaco-Paraná basin. They crop contains more highly deformed units than the regions
out at the edge of the eastern Andes between 22˚ and 27˚ along the eastern border. The layered Schists belt, charac-
S lat. This belt of outcrops widens southward in the terized by intense tectonic layering that was later folded
Pampean ranges, forming a broad block-faulted on a microscale, grades westward into the Gneiss belt
foreland. The Pampean ranges are structurally separated (Figure 4). Postkinematic garnets in the layered Schists
from the Andean chain. From 22˚ to 26˚ S lat, the Precam- belt give 580–540 Ma ages (Bachmann and Grauert,
brian crystalline basement is covered by Cambrian and 1987). The western margin of the crystalline core is made
Ordovician marine successions. To the east, it plunges up of intensely deformed rocks. Mylonite belts along the
below the early Paleozoic succession. These deposits eastern edge of the Puna and at El Peñón can be corre-
partially fill the Chaco plain and the Chaco-Paraná basin lated with the Fiambalá complex (1200 Ma) (Villar and
(Fernández Garrasino, 1989). South of 24˚ S lat, the Coleman, 1986). Between these two mylonite belts, a
western edge of the Precambrian basement is tectonically low-grade schist belt crops out (the Laguna Blanca belt).
related to the Cambrian and Ordovician units. Neverthe- These rocks show a completely different structure to
less, it is also unconformably overlain by these deposits those of the neighboring units (Figure 6).
south of 27˚ S lat (Turner, 1967). Some isolated polydeformed schists crop out away
The relationships among the northern Argentina crys- from the crystalline core (Arequipa-Antofalla belt) and
talline basement, the Brazilian shield, the Precambrian are separated from it by the lower Paleozoic Ocloyic belt
basement in northern Chile, and the Arequipa craton are (Figures 4, 5). They can also be assigned to the Precam-
not known (Figure 5). The most westward outcrops of brian basement (Segerstrom and Turner, 1972). These
the Puna, separate from the crystalline core, seem to be schists are intruded by basic dikes (Allmendinger et al.,
continuous with the northern Chilean outcrops. 1982).
Polydeformation is the main structural characteristic The numerous granitic plutons intruded in the central
of the northern Argentine basement where multiple crystalline basement record a wide range of origins and
episodes of folding and metamorphism have succes- ages (Mon and Hongn, 1991). The structural relation-
sively overprinted one another. The folds show a ships among the distinct Precambrian belts and the
predominant eastward vergence. Considering the entire granitoid intrusives are schematically shown in the inter-
crystalline basement, the degree of metamorphism and pretative cross section of Figure 6, which corresponds to
deformation increases to the west. The metamorphic Precambrian time.
belts have been intruded by several generations of
granitic rocks. Although the age of the crystalline
basement is constrained by the Cambrian and Ordovi-
cian deposits that cover its flanks and its northern SEDIMENTARY AND TECTONIC
margin, radiometric dating shows considerable scatter. EVOLUTION OF PALEOZOIC BASINS
Ages range from 1500 Ma to late Paleozoic (Cingolani
and Varela, 1975; Bahlburg and Breitkreuz, 1991; Rapela Cambrian and Ordovician
et al., 1992; Toselli et al., 1992). The significance of the
isotopic data younger than 600 Ma is uncertain, and The lower Paleozoic sedimentary evolution of the
these dates probably reflect isotopic rejuvenation region began with the deposition of a 2000-m-thick
(Bachmann and Grauert, 1987; Mon and Hongn, 1991). Cambrian quartzite (Mesón Group) (Turner, 1960). This
The most recent radiometric data seems to confirm ages crops out as a narrow belt in the present-day Eastern
older than 1000 Ma (Varela and Dalla Salda, 1992; Dalla Cordillera (Figures 3, 6). Deposition is believed to have
Salda et al., 1993; Demange et al., 1993; Mc Donough et taken place in an elongate north-south striking rift
al., 1993). (Salfity et al., 1975). The Mesón Group covers the low-
Figure 3—Geologic map of northwestern Argentina. Key: 1, Precambrian; 2, Precambrian and Phanerozoic granitoids;
3, Cambrian; 4, Ordovician; 5, Silurian–Devonian; 6, upper Paleozoic; 7, Cretaceous; 8, Tertiary sediments; 9,
Tertiary–Quaternary volcanics; 10, Quaternary sediments; 11, salar; 12, thrusts.
Tectonic Evolution of the Andes of Northern Argentina 273

Figure 4—Regional distribution of the Precambrian belts in


the northern Argentine Andes. Key: a, Precambrian–lower
Paleozoic metamorphics and granitoids; b, thrusts;
c, salar.

grade metamorphic units and its intrusive rocks with a


marked angular unconformity in the eastern border of
the Precambrian core. The rift marks the beginning of a
lengthy stage of extensional tectonism, lasting from the
Late Cambrian to the Late Ordovician (Ocloyic orogeny).
The original shape of the Cambrian rift is not well
known. Despite this, a clastic source in the southeast has
been deduced (Kumpa and Sánchez, 1988). The Precam-
brian crystalline basement was starting to emerge at the
time (Central craton) (Bracaccini, 1960). The present-day
Puna, forming the western flank of the rift, probably Figure 5—Regional framework of the Ocloyic belt in the
provided some of the sediments from a rift flank uplift. Central Andes of Argentina and neighboring regions. Key:
The southern end of the Cambrian rift coincides with the 1, covered Arequipa craton; 2, areas with penetrative Pan-
American deformation; 3, areas with Ocloyic deformation;
present northwest-striking El Toro lineament that 4, covered Ocloyic foldbelt; 5, lower Paleozoic areas with
records a protracted structural history of strike-slip no Ocloyic deformation; 6, Brazilian shield.
movements (Salfity, 1985).
After a period of denudation and tilting related to the
Iruyic movements (Figures 2, 6), the ancient Cambrian turbidites (Turner and Méndez, 1975; Moya, 1988;
rift widened and deepened. Between latest Cambrian Bahlburg, 1990). Ordovician sedimentation occurred in a
and early Llanvirn time, it developed an extensional complex setting that has resulted in a variety of interpre-
basin and a thick succession of platform sediments and tations and regional geodynamic models (e.g., see
Figure 6—Diagram of Precambrian belts, Phanerozoic sedimentary cycles, and tectonic events in northwestern Argentina.
Not to scale. The sections were drawn along 23˚ and 24˚ S lat. Palinspastic restoration not given. Dotted pattern, sedimentary
fill; crosses, granitoids; V, volcanics.
Tectonic Evolution of the Andes of Northern Argentina 275

Bahlburg, 1990). However, Ordovician sedimentation


occurred in two different basins separated by a topo-
graphic high called the Cobres high (Moya, 1988) (Figure
6). The Cobres high became active after the Iruyic
movements. From latest Cambrian to Tremadoc, Arenig,
and early Llanvirn time, more than 4000 m of platform
sediments were deposited in the eastern basin. This basin
coincides with the present-day Eastern Cordillera and
shows no record of synsedimentary magmatic activity.
In the present-day Puna, the western basin was
deeper and more active. A middle Arenig–early Caradoc
succession of turbidites is more than 7000 m thick
(Bahlburg, 1990). In the western Puna, the lower part of
this sequence contains abundant volcaniclastics. These
volcanics suggest intense magmatic activity during the
middle–late Arenig (Coira, 1973; Bahlburg, 1990; Coira
and Koukharsky, 1991). Sedimentation is attributed to an
extensional basin between a volcanic arc forming its
western margin and the Cobres high (Figure 6). High
subsidence rates are compatible with a back-arc basin
setting (Bahlburg, 1990). The base of this rift succession is
not observed on the Puna. Only fossiliferous Tremadoc
beds have recently been reported in the western Puna
(Moya et al., 1993). The locus of sedimentation would
have migrated from east to west. Subsidence is also more
pronounced in that direction, as shown by the thick
turbidite sequences that crop out in the Puna basin.

Ocloyic Foldbelt
The Ocloyic foldbelt of the central Andes in northern Figure 7—Present-day location of the main post-Ocloyic
Argentina (Mon and Hongn, 1991) abuts the western depocenters. Isopach lines are in kilometers. Key: 1,
flank of the crystalline core (Figures 3, 4, 6). This foldbelt Silurian–Devonian; 2, late Paleozoic; 3, Triassic; 4, Creta-
consists of Ordovician deposits covered by a discordant ceous and Tertiary; 5, Tertiary; 6, northern edge of the
Silurian–Devonian succession. Its structure and strati- Chaco-Paraná basin; 7, wedgeout of the Serra Geral lava
graphy clearly distinguish it from the Precambrian crys- flows; 8, faults and lineaments; 9, depocenter; 10, hydro-
carbon field.
talline basement. The Ocloyic foldbelt contains pre-
tectonic granites, assigned a Late Ordovician age. The
Ocloyic deformation is especially visible in the Ordovi-
cian marine sequence of the western Puna basin. separated the Chaco-Paraná basin from the Andean
North of 24˚ S lat, in the Eastern Cordillera, unde- region from at least Early Silurian time onward. North of
formed Cambrian–Ordovician deposits lie with angular 22˚ S lat, where the Ocloyic orogeny was less intense, this
unconformity on the Precambrian crystalline basement positive element does not reach the surface.
(South American para-autochthonous). In the Ordovi-
cian beds, deformation increases gradually to the west Silurian–Devonian Basin
toward the Puna. The deformation ranges from less
deformed sequences to beds affected by intense west- The intracontinental Silurian–Devonian basin of the
verging folding with east-dipping axial plane cleavage. Peruvian and Bolivian Andes extended into northern
Folding did not affect the Precambrian crystalline Argentina. The Argentine part of the basin consists of
basement where the Ocloyic foldbelt is replaced by a two depocenters separated by the Puna arch (present-
west-verging fault belt. The Ocloyic foldbelt shows a day Eastern Cordillera) and the Central craton (Figure 6).
simple folding style, with folds having a single axial The eastern basin coincides with the present-day sub-
plane cleavage surface. Andean ranges and the Chaco plain (Figure 7), while the
South of 24˚ S lat, the contact between the Ocloyic western basin coincides with the Puna. These were intra-
foldbelt and the Precambrian crystalline basement is continental basins, the eastern one containing thick
tectonic. The Precambrian crystalline basement and its platform deposits. This foreland basin corresponds to an
intruding granitoids clearly ride over the Ocloyic foldbelt extensive half-graben fill that thins toward the east. To
(Mon and Hongn, 1988, 1991) (Figure 4). The Ocloyic the west, these deposits are bounded by thrusts that have
orogeny produced the uplift of the Precambrian crys- marked the eastern edge of the Central craton and
talline basement between 22˚ and 34˚ S lat. This elevation Eastern Cordillera probably since Ordovician time
generated a high (Central craton) (Bracaccini, 1960) that (Baldis et al., 1976). Over the western side of the basin,
276 Mon and Salfity

deposition was more than 3000 m thick at the northern Inversion tectonics of the late Paleozoic basins
end. The deposits gradually thin toward the east and probably coincide with the inter-Permian movements
south (Figure 7). The southern end of the present basin (San Rafael orogeny) (Figure 2). The inversion is only
coincides with the northeast-southwest striking Charata recorded in the western Pampean ranges within the
arch (Padula et al., 1967). Other structural highs associ- Paganzo basin realm (Salfity and Gorustovich, 1983)
ated with similar striking faults, such as the Michicola (Figure 7).
and Quirquincho arches, bound more modern depocen-
ters (Figure 7).
The Silurian–Devonian deposits of the western Puna CRETACEOUS–EOCENE AND NEOGENE
basin (Aceñolaza et al., 1972) are represented by isolated
outcrops beneath upper Paleozoic rocks. These beds are SUBSIDENCE
up to 120 m thick and were deposited in a shallow The tectonic and sedimentary evolution of the Creta-
marine environment (Donato and Vergani, 1985). They ceous continental basins of northern Argentina are
discordantly overlie folded Ordovician beds of the reflected in the Salta Group rift. The pre-Cretaceous
Ocloyic foldbelt. These deposits correspond to the basement is heterogeneous and consists of Precambrian,
eastern border of the platform basin that extended into Cambrian, Ordovician, Silurian, Devonian, and
northern Chile, where thicknesses of more than 2500 m Carboniferous units. This basement underwent strong
are recorded (Bahlburg et al., 1987) (Figure 6). Farther subsidence during the first stage of synrift sedimentation
west, in the Coastal Cordillera, they consist of a thick when the former Ocloyic and Chañic structures were
Upper Devonian–Carboniferous turbidite succession reactivated (Salfity, 1979).
deposited in a narrow intracontinental trough (Bahlburg At the beginning of the Cretaceous, a series of grabens
et al., 1987; Niemayer, 1989). Sedimentation appears to opened in northern Argentina, at first as isolated
have migrated from east to west, following increasingly depocenters that were later connected. The most
intense subsidence in that direction. prominent of these depocenters are the Lomas de
Olmedo, Tres Cruces, Sey, Alemanía, and Metán troughs
Chañic Belt (Reyes, 1972; Salfity, 1980; Schwab, 1984) (Figure 7).
The southern limits of this tract of basins were the
The Late Devonian–Early Carboniferous Chañic Traspampean and Pampean arches (Padula and
orogeny is represented by a regional-scale unconformity Mingramm, 1963). The Lomas de Olmedo trough
(Figures 2, 6). This orogeny caused the reversal of the developed between the Michicola and Quirquincho
Silurian–Devonian foreland basin. Chañic deformation arches (Vilela, 1965; Salfity, 1982). The basin also
was also associated with tilting in the Chaco-Paraná contained internal structural highs such as San Pablo and
basin, where the Carboniferous lies unconformably on Salta-Jujuy (Reyes, 1972) (Figures 6, 7).
the Devonian and Precambrian. In the Argentine Puna, The rift depocenters follow three main structural
the Carboniferous also lies discordantly on the trends: north-south, northwest-southeast, and northeast-
Silurian–Devonian succession. southwest. The latter is the most important and is the
In the Chilean coastal Cordillera belt, Upper result of reactivation of the Chañic structures (Bianucci et
Devonian–Carboniferous turbidites were intensely al., 1984; Salfity, 1985). The Kimmeridgian(?)–Campanian
folded during the Toco orogeny (Bahlburg et al., 1987). synrift red beds of the Pirgua Subgroup filled the basins.
This orogeny is considered to be coeval to or slightly Cretaceous extension is believed to have started at the
younger than the movements of the Chañic orogeny. The time of the Araucan movements and was accompanied
Toco orogeny produced south- and southwest-vergent by alkaline volcanism and plutonism (Figures 2, 6). The
folding, with east- and northeast-dipping axial plane postrift sediments were characterized by sandstones and
cleavage (Bell, 1987; Bahlburg et al., 1987). limestones (Yacoraite Formation) in a restricted
carbonate basin (Marquillas, 1985). The final filling of the
basin occurred during the Eocene, when lacustrine and
Tectonic Framework of Upper Paleozoic fluvial deposits accumulated.
Deposits Internally, the Salta Group basins record no regional
angular unconformities. The sedimentary fill was punc-
The Carboniferous (continental) and Permian tuated only by local or regional discontinuities reflecting
(marine) deposits are distributed on either side of the sedimentary or volcanic processes.
Puna arch (Figures 6, 7). In the Tarija and Chaco-Paraná
basins, the Carboniferous is represented by diamictite,
shale, and sandstone sequences more than 1000 m thick
Tectonics
(Mingramm et al., 1979). In the western Puna basin, the The Salta Group basins underwent uplift and erosion
continental Carboniferous succession is about 200 m on at least three occasions after it was completely filled
thick, and farther south in the Paganzo basin it is more (Salfity and Marquillas, 1994) as a result of the following
than 2000 m thick (Figure 7). In the Tarija and Puna events: (1) Inca movements during the late Eocene–early
basins, the Carboniferous lies on Silurian–Devonian Oligocene, (2) the second pulse of the Quechua orogeny
deposits. In the Chaco-Paraná and Paganzo basins, it lies in the late Miocene, and (3) the Diaguita orogeny in the
on Precambrian, Ordovician, and Devonian deposits. Pleistocene (Figures 2, 6).
Tectonic Evolution of the Andes of Northern Argentina 277

During the Inca movements, the Salta Group basins Neogene basins, reactivated the ancient faults bounding
started to emerge, interrupting sedimentation and the Salta Group basins. These faults began the reverse slip
resulting in erosion and stripping of several different during the Quechua orogeny. Some normal faults that
formations of the postrift stage. These movements were bounded the Cretaceous Pirgua Subgroup graben became
generally basinwide, although structurally they were not west-vergent thrusts in the central sub-Andean ranges,
intense. The erosion was recorded along the western between 24˚ and 25˚ S lat (Figure 3). Similarly, Cretaceous
flank of the Pampean arch (Salfity et al., 1993). After the east-west striking south-vergent normal faults reactivated
Incaic event, many of the Salta Group basins remained by the Diaguita orogeny are known in the Calchaquí
buried under Paleogene and Neogene deposits. Sedimen- Valley at 26˚ S lat (Grier et al., 1991) (Figure 3).
tation occurred from late Oligocene (Pehuenche
movements) time onward. The deposition took place
within a foreland basin in the sub-Andean ranges and CENOZOIC STRUCTURES
within intermontane basins in the Puna region (Figure 6).
The Neogene continental sequences of northern The Cenozoic Andean chain was formed through
Argentina contain no contemporaneous volcanic units in interaction of the Nazca plate with the South American
their basal sections. This reflects a period of magmatic plate. This occurred from Miocene time onward (Dewey
calm between the Eocene and the middle Miocene that is and Bird, 1970; Jordan et al., 1983). Within this sector of
characteristic of northern Argentina (Moya and Salfity, the eastern Andean slope there are belts of varying struc-
1982). tural style. The most westerly of these corresponds to the
The second episode of the late Miocene Quechua Altiplano or Puna. This plateau has an average elevation
orogeny produced folding and erosion of the Salta of 4000 m above sea level with internal drainage feeding
Group formations in several parts of the basin. The enclosed depressions and salt flats. It is formed by
upper part of the Tertiary conglomeratic interval Cenozoic volcanic arc and plates of Paleozoic rocks
contains boulders from the Yacoraite limestone and other thrust eastward. Among them, remnants of the Creta-
Salta Group units (Russo and Serraioto, 1979). They ceous and Tertiary continental cover are found. South of
consist of thick psephitic successions deposited in 24˚ S lat, the eastern edge of the Puna is made up of
taphrogenic basins. Precambrian basement sheets thrust eastward over the
The profuse volcanic activity is attributed to the Cretaceous and Tertiary continental successions. These
Miocene–Pleistocene ensialic magmatic arc that started to deposits fill the Calchaquí Valley depression. North of
develop on the western flank of the Puna (Figures 1, 3, 6). 24˚ S, the Puna is not clearly separated from the Eastern
The host rocks of this volcanic arc are believed to be Cordillera. The structure of the Puna at depth is shown
Paleozoic or Precambrian rocks that make up the schematically in Figure 8e.
Andean arch (Salfity and Marquillas, in press) (Figure 6). The Calchaquí Valley forms an elongate north-south
The Andean arch approximately coincides with the depression. From 24˚ to 26˚ S lat, it extends between the
Huaytiquina high and developed along the western Puna and the Eastern Cordillera and farther south into
flanks of the San Pablo and Traspampean arches (Figures the Pampean ranges. The valley is bounded by reverse
6, 7). All the highs are composed of Precambrian or faults that display definite horizontal components,
Paleozoic rocks. Neither the Cretaceous rocks nor the commonly set en echelon, indicating a transpressive
basal part of the Tertiary sequence of the Uyuni, Puri- setting associated with sinistral rotation. The vergence of
lactis, Arizaro, and Antofalla basins (Figures 6, 7) hosted the faults along each border is toward the axis of the
the volcanic arc. depression. The edges of the Calchaquí Valley are largely
The conspicuous reversal of the Salta Group basins made up of Precambrian basement rocks. Folded Creta-
occurred during the compressive Diaguita orogeny. This ceous and Tertiary sedimentary rocks crop out in the
orogeny was the most important of the Andean depression.
movements (Figure 2), which folded and faulted the The Eastern Cordillera is a fold and thrust belt
Cretaceous and Cenozoic sedimentary columns and following the eastern side of the Puna. South of 24˚ S lat,
commonly reactivated the pre-Cretaceous basement. the Calchaquí Valley separates these two geologic
Their effects reached eastward as far as 64˚ W long provinces. The highest ranges in the Eastern Cordillera
(Figure 3), eastward of which the Salta Group basins reach altitudes above 5000 m. The Eastern Cordillera
remain covered by the Orán Group deposits which are consists mainly of sheets of faulted Precambrian
about 3000 m thick (Figure 6). These movements did not basement, but it shows marked along-strike differences
attain the north Puna. in its stratigraphy. North of 25˚ S lat, it contains thick
The effects of the Diaguita orogeny were particularly Cambrian and Ordovician strata. Between 25˚ and 26˚ S
pronounced. At present, lacustrine deposits of the lat, the Paleozoic successions are replaced by Creta-
Yacoraite Formation are found 4500 m above sea level, ceous–Tertiary continental rocks that have a combined
100 km to the west-northwest of Salta. Furthermore, the thickness greater than 4000 m. The Tertiary deposits lie
oil wells, drilled 160 km to the northeast of Salta in the on Precambrian basement south of 26˚ S lat due to the
southern Lomas de Olmedo depocenter, have penetrated absence of Mesozoic and Paleozoic rocks (Figure 7).
the top of the Yacoraite Formation at 6100 m below the North of 24˚ S lat, the Eastern Cordillera is a thrust belt
surface. The Diaguita orogeny, which involved pre- with eastward vergence and a typical piggy-back struc-
Cretaceous basement, Salta Group deposits, and tural style. The thrust displacements of Precambrian
278
Mon and Salfity

Figure 8—Regional cross sections of northwestern Argentina. See locations in Figure 3. Key: (A) Transitional contact between the Puna and Eastern Cordillera
and the deformed Cambrian graben. (B) Thin-skinned northern sub-Andean ranges (after Aramayo Flores, 1989). (C) Eastern Cordillera thrust over thick-skinned
sub-Andean ranges and the west-verging Santa Bárbara System. (D) Thick-skinned west-verging Santa Bárbara System. (E) Southern Puna, Precambrian
basement thrust over Ordovician deposits, eastern edge of the Puna and Calchaquí Valley. Key: 1, Precambrian; 2, Precambrian and Phanerozoic granitoids;
3, Cambrian; 4, Ordovician; 5, Silurian–Devonian; 6, late Paleozoic; 7, Cretaceous; 8, Tertiary sediments; 9, Tertiary–Quaternary volcanics; 10, Quaternary
sediments.
Tectonic Evolution of the Andes of Northern Argentina 279

basement sheets over younger deposits are more than 10 foreland basin (Figures 6, 7) consists of three transgres-
km (Figure 8a). Southward, the basement blocks on the sive–regressive cycles (Aramayo Flores, 1989; Vistali,
western edge of the Eastern Cordillera are thrust 1989). The Devonian Los Monos Formation is the source
westward (3–4 km displacement) over the Cretaceous– rock and the Santa Rosa Formation is the reservoir
Tertiary successions of the Calchaquí Valley (Figures 3, formed by secondary fracturing (Figure 2). The Devonian
8e). In this sector, the Eastern Cordillera displays reservoirs occur at a depth of 4000 m.
eastward and westward vergence. The back-thrusting is Devonian, Carboniferous, and Tertiary sedimentary
probably related to deep-seated blind thrusts, as is successions are superimposed in the area of the present-
observed in other mountain chains where the subsurface day fold and thrust belt. Gas has been found in Devonian
is better known (e.g., the Variscan belt of Europe, G. reservoirs of the San Antonio and Aguaragüe ranges
Drozdzewski, 1992, personal communication). (Figure 8b) (Aramayo Flores, 1989). Hydrocarbon reser-
The eastern edge of the Eastern Cordillera is marked voirs also occur in the intracratonic Carboniferous Tarija
by a large regional thrust separating it from the sub- basin, as well as in the superimposed Tertiary foreland
Andean ranges (Mon, 1991). The displacement on this basin (Figures 2, 6). Devonian oil from the Los Monos
thrust is as much as 15 km (Figures 3, 8c). Formation migrated into the Carboniferous Tupambi
The sub-Andean ranges form part of the outermost and Tarija formations and Tertiary lower Orán Group
belt of the Andean foreland, as far south as 27˚ S lat. The reservoirs (Aramayo Flores, 1989).
region consists of a foldbelt made up of three large anti- South of the sub-Andean ranges, the Carboniferous
clinoria that are separated by axial depressions. Between successions overlie Devonian strata in the Alhuampa
24˚ and 25˚ S lat, the central anticlinorium shows a basin (Figure 7). No hydrocarbon discoveries have yet
westward vergence with back-thrusts on the western been made in this basin. The Michicola, Quirquincho,
flank of the folds (Figure 8c). As on the western edge of and Pampean arches separate the present Tarija and
the Eastern Cordillera, these structures are probably Alhuampa basins. Commercial hydrocarbons have also
related to subsurface blind thrusts. The structure of the not been found in the continental Carboniferous–
northern segment of the sub-Andean ranges is related to Permian and Triassic successions of the Paganzo basin
basal décollement linked to thin-skinned thrusts (Figure 7).
(Mingramm et al., 1979; Aramayo Flores, 1989) (Figure The intracontinental Cretaceous basins of the region
8b). The remainder of the belt is influenced by substantial are productive mostly from the Lomas de Olmedo
basement involvement in sub-Andean folding (Figures depocenter; there is only one field in the Metán
8c, 8d). depocenter (Figure 7). The sedimentary fill of the Lomas
South of 24˚ S lat, the San Francisco and Juramento de Olmedo depocenter, as well as the other Salta Group
rivers occupy a depression between the thrusts which basins, consists of prerift and synrift continental red beds
marks the eastern border of the Eastern Cordillera and and volcanics of the Pirgua Subgroup (Figure 2) (Salfity
the western heights of the sub-Andean ranges (Figure 3). and Marquillas, 1994). Postrift deposition took place
This depression contains almost 10-km-thick Cretaceous during the latest Cretaceous–Paleogene. The Yacoraite
and Tertiary deposits. The deep structure of this depres- limestone is the most conspicuous stratigraphic unit
sion is not known in detail. However, reconstructions among the postrift formations (Figure 2). This formation
based on surface data suggest a complex structure forms source rock, reservoir, and seal.
(Figures 8c, 8d). The tectonic origin of the Lomas de Olmedo
At 27˚ S lat, the Andean foreland undergoes a pro- depocenter is related to the extensive Araucan exten-
nounced change in structural style (Mon, 1976; Jordan et sional movements along the Central Andes during Late
al., 1983). The sub-Andean ranges are replaced south- Jurassic–Early Cretaceous time (Figure 2). The area
ward by the suite of Precambrian basement blocks of the occupied by the large Cretaceous–Tertiary Lomas de
Pampean ranges, with dominantly west-vergent Olmedo depocenter was originally a structural high,
thrusting. These probably developed as back-thrusts due bounded by northeast-southwest lineaments, probably
to the action of blind thrusts at depth. The most westerly from Late Devonian–Early Carboniferous time onward
blocks are eastward vergent and define an axis of (Figure 7). This high was part of the Michicola arch
tectonic convergence that is the southern continuity of during the time of filling of the Carboniferous Tarija
those observed in the Calchaquí Valley. basin (Salfity et al., 1987).
Erosion occurred in the Lomas de Olmedo area,
affecting the pre-Cretaceous basement from the Late
HYDROCARBON HABITAT Devonian–Early Carboniferous onward. The erosion was
most intense along the northeast-southwest axis of the
The Silurian–Devonian and Cretaceous basins Lomas de Olmedo depocenter (Figure 7). The Silurian–
produce oil and gas in the sub-Andean ranges and in the Devonian succession was partially or totally eroded, so
Chaco plain (Figure 7). Hydrocarbons in Carboniferous that the pre-Cretaceous basement consists of Ordovician,
and Tertiary reservoirs were derived from Devonian Silurian, and Devonian strata. Erosion is believed to have
source rocks. Other basins such as the Alhuampa stripped much of the Silurian–Devonian succession as a
(Chaco-Paraná) and Paganzo (Bolsones) basins are still consequence of which Ordovician rocks directly underlie
being explored. the Salta Group (Padula et al., 1967; Mingramm and
The clastic fill of the post-Ocloyic Silurian–Devonian Russo, 1972; Salfity, 1979; Carlé et al., 1989).
280 Mon and Salfity

The strip along which the greatest amount of erosion Matte and Xu Zhi (1988) for the Variscan belts of central
occurred coincides with the Late Jurassic–Early Creta- Europe and Asia. The terrane situated west of the
ceous (Araucan extensional event) to late Tertiary Ocloyic belt may represent the southern continuation of
depoaxis. The greatest thickness of Cretaceous (Salta the Arequipa craton, which has remained close to South
Group) and Tertiary (Orán Group) successions in the America since 850 Ma (Shackleton et al., 1979; Baeza and
Lomas de Olmedo depocenter is at least 6 km. This Pichowiak, 1988) (Figure 5).
depocenter thus reflects tectonic behavior and erosion The synsedimentary volcanic rocks, the plutons
from late Paleozoic time onward. intruded in the Ordovician beds, and the plutons
The time of oil generation in the Devonian and Creta- intruded along the western border of the Precambrian
ceous basins is believed to have been related to Neogene crystalline basement core had originally been assigned to
burial history. The time of oil migration in the Tarija the eastern Puna eruptive belt (Méndez et al., 1973). Later
basin took place in the late Neogene (Jordan and Alonso, studies (Coira et al., 1982) ascribed them to the roots of
1987). It is interesting that, almost 50 years ago, Reed an Ordovician magmatic arc associated with the
(1946) observed the great difference in age between evolution of the Ocloyic belt. This hypothesis has been
Devonian source rock deposition and late Tertiary trap incorporated in most of the geotectonic models of the
formation in the Tarija basin. He concluded that hydro- region (e.g., Bahlburg, 1990; Dalla Salda et al., 1992a,
carbon generation and migration was driven by Tertiary 1992b). According to Mon and Hongn (1991), the
deformation. granites of the western border of the Precambrian crys-
The oil in the Cretaceous basin is also believed to have talline basement are separated from the Ocloyic foldbelt
been recently generated. Generation probably began at and its magmatic components by a large Ocloyic thrust.
the end of Orán Group deposition (late? Oligocene to Moreover, along the western border of the Precambrian
later Pliocene–early Pleistocene). These sediments cover basement, there are three generations of granitoids. The
the Lomas de Olmedo and Metán depocenters (Figure 7). older plutons are affected by ductile shear belts probably
The optimum overburden thickness for oil genesis was of Precambrian age. Therefore, it is difficult to reconcile
4000–5000 m in the Lomas de Olmedo basin. Migration these characteristics with a single Ordovician magmatic
filled the sandstone reservoirs with primary porosity in arc related to the Ocloyic belt evolution, as has been
the Yacoraite Formation in the Chaco basin and with suggested.
secondary fracture porosity in the Yacoraite limestone in The genesis of the Silurian–Devonian foreland and the
the sub-Andean ranges. Carboniferous intracratonic basins took place after the
Ocloyic orogeny. These basins were affected by vertical
and flexural movements. The Chañic orogeny, approxi-
DISCUSSION mately equivalent to the Toco orogeny of northern Chile,
is represented in northern Argentina by a low-angle
The Precambrian crystalline basement consists of unconformity separating the Carboniferous–Permian
polydeformed metamorphic rocks that were invaded by beds from the Silurian–Devonian succession. The Toco
various generations of granitoids. It is a complex orogeny in the Chilean coastal Cordillera generated a
structure constructed by a variety of units, some of foldbelt with penetrative deformation (Bahlburg et al.,
which are probably allochthonous. This complex was 1987; Bell, 1987). The folding associated with the San
amalgamated by the Pan-American orogeny at the Rafael orogeny in the western border of the Pampean
Precambrian–Cambrian boundary. Since then, the ranges, south of 28˚ S lat, is also represented in the study
basement has suffered little remobilization capable of area by a low-angle unconformity.
changing its internal structure (Mon and Hongn, 1991; Paleozoic tectonic evolution shows a definite
Demange et al., 1993). Paleozoic folding and penetrative westward trend with deformation belts becoming
deformation were restricted to the Late Ordovician– progressively younger in that direction. This behavior
Early Silurian Ocloyic belt and to the Pacific coast Late was suddenly reversed in the latest Paleozoic. From
Carboniferous–Permian Chañic belt. During the greater Jurassic time onward, younger magmatic arcs progres-
part of Phanerozoic time, the crystalline core remained as sively developed to the east along the western slope of the
a positive or nearly positive element. Paleozoic Andes in northern Chile. This intense Mesozoic magmatic
movements resulted in faulting and ductile shear zones activity did not reach northern Argentina. Extensional
along the west edge (H. D. Hongn, 1993, personal subsidence responsible for the Salta basins and associated
communication). magmatism started during the Late Jurassic or Early
The Ocloyic foldbelt evolved between the Precam- Cretaceous (the Araucan event). Extension continued
brian crystalline basement core and a para-autochtho- during Late Cretaceous and Paleogene postrift times.
nous terrane situated farther west represented by the Early Oligocene Inca compression inverted the basin and
Arequipa-Antofalla belt and the crystalline basement interrupted sedimentation. During the Neogene, conti-
outcrops of northern Chile (Figures 4, 5). Both terranes nental sedimentation took place in the sub-Andean
collided along an eastward-dipping continental suture. foreland basin and in extensional intermontane basins of
The folding of the Ocloyic belt may be related to a basal the Puna.
décollement, which does not attain the surface, allowing The Neogene tectonic evolution of this part of the
the sliding of the Ordovician succession over the crys- Andes took place over a segment of the subducting
talline basement (Figure 6). This model was proposed by oceanic Nazca plate that dipped eastward at 30˚ beneath
Tectonic Evolution of the Andes of Northern Argentina 281

the continental margin of South America (Jordan et al., Allmendinger, R. W., T. Gubbels, B. Isacks, and T. Cladouhos,
1983). The first important Neogene deformation was the 1993, Lateral variations in late Cenozoic deformation,
Quechua orogeny, which occurred at about 10 Ma (Jordan central Andes, 20˚–28˚ S: Second International Symposium
and Alonso, 1987). This orogeny affected the Puna and on Andean Geodynamics, Expanded Abstracts, p. 155–158.
Aramayo Flores, F. R., 1989, El cinturón plegado y sobrecor-
uplifted the Eastern Cordillera, deforming the western
rido del norte argentino: Boletín de Informaciones Petrol-
flank of the Salta Group basins. The Diaguita orogeny, eras, Tercera Epoca, no. 17, p. 2–16.
younger than 3–4 Ma, was according to Jordan and Bachmann, C., and G. Grauert, 1987, Datación de metamor-
Alonso (1987) the most important tectonic episode of fismo basada en el análisis isotópico Rb/Sr en perfiles de
Neogene Andean evolution. It generated most of the pequeña sección de metasedimentos polimetamórficos en
eastward-verging thrusts, inverted the Mesozoic and el noroeste argentino: Décimo Congreso Geológico
Cenozoic basins, and reactivated the west-verging Ocloyic Argentino, Actas 3, p. 17–20.
thrusts in the Eastern Cordillera and Pampean ranges. It Baeza, L., and S. Pichowiak, 1988, Ancient crystalline
was responsible for the thrusting of the Puna over the basement provinces in the northern Chilean central
Pliocene–Pleistocene conglomerates, which were pro- Andes—relicts of continental crust development since the
mid Proterozoic, in H. Bahlburg, C. Breitkreuz, and P.
duced by erosion of its eastern flanks. The effects of the
Giese, eds., The southern Central Andes: Lecture Notes in
Diaguita orogeny extended farther east to the sub-Andean Earth Sciences, Berlin, Springer–Verlag, v. 17, p. 3–24.
and Pampean ranges. According to Jordan and Alonso Bahlburg, H., 1990, The Ordovician basin in the Puna of NW
(1987) and Allmendinger et al. (1993), the Diaguita Argentina and N Chile: geodynamic evolution from back-
orogeny did not affect the Puna north of 23˚ S lat. arc to foreland basin: Geotektonische Forschungen, v. 75,
p. 1–107.
Bahlburg, H., and Ch. Breitkreuz, 1991, Paleozoic evolution of
active margin basins in the southern Central Andes (north-
CONCLUSIONS western Argentina and northern Chile): Journal of South
American Earth Sciences, v. 4, p. 171–188.
Northern Argentina is an ideal place to study the rela- Bahlburg, H., Ch. Breitkreuz, and W. Zeil, 1987, Paleozoic
tionships between a young orogenic belt, such as the basin development in northern Chile (21˚–27˚ S): Geolo-
Andes, and the Precambrian crystalline basement. Wide gische Rundschau, v. 76, p. 633–646.
outcrops of the roots of the Andean fold and thrust belt Baldis, B. A., A. Gorroño, J. V. Ploszkiewicz, and R. M. Sarudi-
show that the basement is deeply involved in the young ansky, 1976, Geotectónica de la Cordillera Oriental, Sierras
structures and that Andean folding is not related to a Subandinas y comarcas adyacentes: Sexto Congreso
regional basal décollement. The folding of other Geológico Argentino, Actas 1, p. 3–22.
segments of the Andean chain, such as the Precordillera, Bell, C. M., 1987, The late Paleozoic evolution of the Gond-
wanaland continental margin in northern Chile:
are possibly related to a basal décollement. Geophysics Monograph, v. 40, p. 261–270.
The post-Ocloyic evolution of this region involved Bianucci, H. A., O. Acevedo, and J. R. Vásquez, 1984, La
continental crust at shallow depth, with no development tectónica varíscica en el noroeste argentino y sus estilos
of penetrative structures. Extrusive magmatism predom- estructurales: Noveno Congreso Geológico Argentino,
inated with minor hypabyssal emplacement. There was a Actas 2, p. 25–30.
close relationship among structural architecture, subsi- Bracaccini, O., 1960, Lineamientos principales de la evolución
dence history, structural inversion, and the timing of oil estructural de la República Argentina: Instituto Argentino
generation and migration in Devonian and Cretaceous del Petróleo, Petrotecnia, v. 10, p. 57–69.
producing basins. Carlé, R. J., O. E. Di Persia, and G. A. Olivieri, 1989, Análisis
geológico y petrolero del sector noroeste de la provincia de
Formosa, República Argentina: Primer Congreso Nacional
de Exploración de Hidrocarburos, v. 1, p. 163–185.
Acknowledgments This paper was greatly improved as a Cingolani, C., and R. Varela, 1975, Geocronología rubidio-
estroncio de las rocas ígneas y metamórficas de las sierras
result of the valuable critical review of A. J. Tankard, H. J. Chica y Grande de Córdoba: Segundo Congreso
Welsink, M. A. Uliana, V. A. Ramos, and an anonymous Iberoamericano de Geología Económica, Actas 1, p. 9–35.
reviewer. Our studies were made possible by grants from Coira, B., 1973, Resultados preliminares sobre la petrología
CONICET of Argentina, Salta National University, and del ciclo eruptivo concomitante con la depositación de la
Tucumán National University. Formación Acoite en la zona de Abra Pampa, Jujuy:
Asociación Geológica Argentina Revista, v. 28, p. 85–88.
Coira, B., and M. Koukharsky, 1991, Lavas en almohadilla
ordovícicas en el cordón de Escaya, Puna septentrional,
REFERENCES CITED Argentina: Sexto Congreso Geológico Chileno, Actas 1,
p. 674– 678.
Aceñolaza, F., J. Benedetto, and J. A. Salfity, 1972, El Neopale- Coira, B., J. Davidson, C. Mpodozis, and V. Ramos, 1982,
ozoico de la Puna argentina: su fauna y relación con áreas Tectonic and magmatic evolution of the Andes of northern
vecinas: Anales Academia Brasileira de Ciencias, v. 44, p. Argentina and Chile: Earth Science Reviews, v. 18,
5–20. p. 303–332.
Allmendinger, R. W., T. Jordan, M. Palma, and V. Ramos, Dalla Salda, L., C. Cingolani, and R. Varela, 1992a, Early
1982, Perfil estructural de la Puna catamarqueña (25˚–27˚ Paleozoic orogenic belt of the Andes in southwestern
S), Argentina: Quinto Congreso Latinoamericano de South America: result of Laurentia–Gondwana collision?:
Geología, Actas 1, p. 499–518. Geology, v. 20, p. 617–620.
282 Mon and Salfity

Dalla Salda, L., I. W. D. Dalziel, C. A. Cingolani, and R. regional argentina: Academia Nacional de Ciencias de
Varela, 1992b, Did the Taconic Appalachians continue into Córdoba, v. 1, p. 95–138.
southern South America?: Geology, v. 20, p. 1059–1062. Mon, R., 1976, La tectónica del borde oriental andino en las
Dalla Salda, L., R. Varela, and C. Cingolani, 1993, Sobre la provincias de Salta, Tucumán y Catamarca: Asociación
colisión de Laurentia–Sudamérica y el orógeno Fama- Geológica Argentina Revista, v. 31, p. 65–72.
tiniano: Décimo Segundo Congreso Geológico Argentino y Mon, R., 1991, Estructura profunda de la cadena subandina
Segundo Congreso de Exploración de Hidrocarburos, entre los 24˚ 30’ y 27˚ 00’ S: Sexto Congreso Geológico
Actas 3, p. 358– 366. Chileno, Actas 1, p. 481–484.
Demange, M., E. G. Baldo, and R. D. Martino, 1993, Structural Mon, R., and F. D. Hongn, 1988, El corrimiento del borde occi-
evolution of the sierras de Córdoba, Argentina: Second dental del Cratógeno central en la Puna: Asociación
International Symposium Andean Geodynamics, Geológica Argentina Revista, v. 43, p. 338–342.
Extended Abstracts, p. 513–515. Mon, R., and F. D. Hongn, 1991, The structure of the Precam-
Dewey, J. M., and J. M. Bird, 1970, Mountain belts and new brian and lower Paleozoic basement of the Central Andes
global tectonics: Journal of Geophysical Research, v. 75, between 22˚ and 32˚ S. latitude: Geologische Rundschau,
p. 2625– 2647. v. 80, p. 745–758.
Donato, E., and G. Vergani, 1985, Geología del Devónico y Moya, M. C., 1988, Lower Ordovician in the southern part of
Neopaleozoico de la zona sur del Cerro Rincón, provincia the Argentine Eastern Cordillera, in H. Bahlburg, Ch.
de Salta, Argentina: Cuarto Congreso Geológico Chileno, Breitkreuz, and P. Giese, eds., The southern Central Andes:
Actas 1, p. 262–284. Lecture Notes in Earth Sciences, Berlin, Springer–Verlag, v.
Fernández Garrasino, C., 1989, La cuenca Chaco–Paranense 17, p. 55–70.
Argentina. Sus tendencias evolutivas y algunas posibili- Moya, M. C., and J. A. Salfity, 1982, Los ciclos magmáticos en
dades exploratorias: Boletín de Informaciones Petroleras, el noroeste argentino: Quinto Congreso Latinoamericano
Tercera Epoca, no. 18, p. 2–17, de Geología, Actas 3, p. 523–536.
Grier, M., J. A. Salfiy, and R. W. Allmendinger, 1991, Andean Moya, M. C., S. Malanca, F. D. Hongn, and H. Bahlburg, 1993,
reactivation of the Cretaceous Salta rift, northwestern El Tremadoc temprano en la Puna occidental argentina:
Argentina: Journal of South American Earth Sciences, v. 4, Décimo Segundo Congreso Geológico Argentino y
p. 351–372. Segundo Congreso de Exploración de Hidrocarburos,
Halpern, M., and C. O. Latorre, 1973, Estudio geocronológico Actas 2, p. 20–30.
inicial de rocas del noroeste de la República Argentina: Niemayer, H., 1989, El complejo ígneo-sedimentario del
Asociación Geológica Argentina Revista, v. 28, p. 195–205. cordón de Lila, región de Antofagasta: significado
Jordan, T. E., and R. N. Alonso, 1987, Cenozoic stratigraphy tectónico: Revista Geológica de Chile, v. 16, p. 163–181.
and basins tectonics of the Andes mountains, 20˚–28˚ south Omarini, R., A. Aparicio, C. Parica, S. Pichowiak, L. García, K.
latitude: AAPG Bulletin, v. 71, p. 49–64. Damm, J. Viramonte, J. Salfity, and R. Alonso, 1985,
Jordan, T. E., B. L. Isacks, R. W. Allmendinger, J. A. Brewer, V. Nuevos datos geocronológicos acerca de la edad precám-
A. Ramos, and J. Ando, 1983, Andean tectonics related to brica de la Formación Puncoviscana, noroeste argentino:
geometry of subducted Nazca plate: GSA Bulletin, v. 94, Comunicaciones, Departamento de Geología de la Univer-
p. 341–361. sidad de Chile, v. 35, p. 181–183.
Kumpa, M., and M. C. Sánchez, 1988, Geology and sedimen- Padula, E., and A. Mingramm, 1963, The fundamental geolog-
tology of the Cambrian Grupo Mesón (NW Argentina), in ical pattern of the Chaco-Parana basin (Argentina) in
H. Bahlburg, C. Breitkreuz, and P. Giese, eds., The relation to its oil possibilities: Sixth World Petroleum
southern Central Andes: Lecture Notes in Earth Sciences, Congress, Frankfurt am Main, Proceedings 1, p. 1–18.
Berlin, Springer–Verlag, v. 17, p. 39–54. Padula, E., E. O. Rolleri, A. Mingramm, P. Criado Roque, M.
Marquillas, R. A., 1985, Estratigrafía, sedimentología y pale- A. Flores, and B. A. Baldis, 1967, Devonian of Argentina:
oambientes de la Formación Yacoraite (Cretácico superior) International Symposium of the Devonian System,
en el tramo austral de la cuenca, norte argentino: Ph.D. Calgary, Proceedings 2, p. 165–199.
dissertation, Universidad Nacional de Salta, Salta, Rapela, C. W., B. Coira, A. Toselli, and J. Saavedra, 1992, El
Argentina, 139 p. magmatismo del Paleozoico inferior en el sudoeste de
Matte, Ph., and Q. Xu Zhi, 1988, Décollements in slate belts, Gondwana, in J. G. Gutiérrez Marco, J. Saavedra, and I.
examples from European Variscides and the Qin Ling belt Rábano, eds., Paleozoico inferior de Ibero-América:
of central China: Geologische Rundschau, v. 77, p. 227–238. Mérida, Universidad de Extremadura, p. 21–68.
Mc Donough, M. R., V. A. Ramos, C. E. Isachsen, S. A. Reed, L. C., 1946, San Pedro oil field, province of Salta,
Bowring, and G. I. Vujovich, 1993, Edades preliminares de northern Argentina: AAPG Bulletin, v. 30, p. 591–605.
circones del basamento de la sierra del Pie de Palo, sierras Reyes, F. C., 1972, Correlaciones en el Cretácico de la cuenca
Pampeanas occidentales de San Juan: sus implicancias andina de Bolivia, Perú y Chile: Yacimientos Petrolíferos
para el supercontinente de Rodinia: Décimo Segundo Fiscales Bolivianos, Revista Técnica, v. 1, p. 101–144.
Congreso Geológico Argentino y Segundo Congreso de Russo, A., and A. Serraioto, 1979, Contribución al
Exploración de Hidrocarburos, Actas 3, p. 340–342. conocimiento de la estratigrafía terciaria en el noroeste
Méndez, V., A. Navarini, D. Plaza, and V. Viera, 1973, Faja argentino: Séptimo Congreso Geológico Argentino, Actas
eruptiva de la Puna oriental: Quinto Congreso Geológico 1, p. 715–730.
Argentino, Actas 4, p. 147–158. Salfity, J. A., 1979, Paleogeología de la cuenca del Grupo Salta
Mingramm A., and A. Russo, 1972, Sierras Subandinas y del norte de Argentina: Sexto Congreso Geológico
Chaco salteño, in A. F. Leanza, ed., Geología regional Argentino, Actas 1, p. 239–255.
argentina: Academia Nacional de Ciencias de Córdoba, Salfity, J. A., 1980, Estratigrafía de la Formación Lecho
p. 185–221. (Cretácico) en la cuenca andina del norte argentino: Ph.D
Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979, dissertation, Universidad Nacional de Salta, Salta,
Sierras Subandinas, in J. C. M. Turner, ed., Geología Argentina, 91 p.
Tectonic Evolution of the Andes of Northern Argentina 283

Salfity, J. A., 1982, Evolución paleogeográfica del Grupo Salta Petroleum basins of South America: AAPG Memoir 62,
(Cretácico–Eogénico), Argentina: Quinto Congreso Lati- this volume.
noamericano de Geología, Actas 1, p. 11–26. Toselli, A. J., L. Dalla Salda, and R. Caminos, 1992, Evolución
Salfity, J. A., 1985, Lineamientos transversales al rumbo metamórfica y tectónica del Paleozoico inferior de
andino en el noroeste argentino: Cuarto Congreso Argentina, in J. G. Gutiérrez Marco, J. Saavedra, and I.
Geológico Chileno, Actas 2, p. 119–227. Rábano, eds., Paleozoico inferior de Ibero-América:
Salfity, J. A., and S. A. Gorustovich, 1983, Paleogeografía del Mérida, Universidad de Extremadura, p. 279–309.
Grupo Paganzo (Paleozoico Superior): Asociación Turner, J. C. M., 1960, Estratigrafía de la sierra de Santa
Geológica Argentina Revista, v. 38. p. 437–453. Victoria y adyacencias: Academia Nacional de Ciencias de
Salfity, J. A., and R. A. Marquillas, 1994, Tectonic and sedi- Córdoba Boletín, v. 41, p. 163–196.
mentary evolution of the Cretaceous–Eocene Salta Group Turner, J. C. M., 1967, Descripción geológica de la hoja 13 b,
basin, Argentina, in J. A. Salfity, ed., Cretaceous Tectonics Chaschuil (provincias de Catamarca y La Rioja): Instituto
of the Andes: Earth Evolution Sciences, Braunschweig, Nacional de Geología y Minería Boletín, v. 106, 79 p.
Fried. Vieweg & Sohn, p. 266–315. Turner, J. C. M., and V. Méndez, 1975, Geología del sector
Salfity, J. A., R. Omarini, B. Baldis, and W. J. Gutiérrez, 1975, oriental de los departamentos de Santa Victoria e Iruya,
Consideraciones sobre la evolución geológica del Precám- provincia de Salta, República Argentina: Academia
brico y Paleozoico del norte argentino: Segundo Congreso Nacional de Ciencias de Córdoba Boletín, v. 51, p. 11–24.
Iberoamericano de Geológica Económica, Actas 4, Varela, R., and L. Dalla Salda, 1992, Geocronología Rb-Sr de
p. 341–361. metamorfitas y granitoides del extremo sur de la sierra del
Salfity, J. A., C. L. Azcuy, O. López G., D. A. Valencio, J. F. Pie de Palo, San Juan: Asociación Geológica Argentina
Vilas, A. Cuerda, and G. Laffitte, 1987, Cuenca Tarija, in S. Revista, v. 47, p. 271–275.
Archangelsky, ed., El Sistema Carbonífero en la República Vilela, C. R., 1965, El petróleo de las cuencas de Orán y Metán:
Argentina: SCCS-Project PICG 211, Academia Nacional de Acta Geológica Lilloana, v. 7, p. 425–438.
Ciencias de Córdoba, p. 15–39. Villar, L., and R. Coleman, 1986, Reinterpretación geológica
Salfity, J. A., C. R. Monaldi, R. A. Marquillas, and R. E. de la faja ultrabásica y el bloque de alto grado de metamor-
González, 1993, La inversión tectónica del umbral de Los fismo de sierra de Fiambalá, provincia de Catamarca:
Gallos en la cuenca del Grupo Salta durante la fase Incaica: Asociación Geológica Argentina Revista, v. 41, p. 410–413.
Décimo Segundo Congreso Geológico Argentino y Vistali, M. C., 1989, La cuenca siluro–devónica del noroeste, in
Segundo Congreso de Exploración de Hidrocarburos, G. A. Chebli, and L. A. Spalletti, eds., Cuencas Sedimenta-
Actas 3, p. 200–210. rias Argentinas: Universidad Nacional de Tucumán, Serie
Schwab, K., 1984, Contribución al conocimiento del sector Correlación Geológica, v. 6, p. 19–41.
occidental de la cuenca sedimentaria del Grupo Salta
(Cretácico–Eogénico) en el noroeste argentino: Noveno
Congreso Geológico Argentino, Actas 1, p. 586–604.
Segerstrom, K., and J. C. M. Turner, 1972, A conspicuous Authors’ Mailing Addresses
flexure in regional structural trend in the Puna of north- R. Mon
western Argentina: USGS Professional Paper 800B, Casilla de Correo 36, Sucursal 2
p. 205–209. 4000 Tucumán
Shackleton, R. M., A. C. Ries, M. P. Coward, and P. R.
Cobbold, 1979, Structure, metamorphism and
Argentina
geochronology of the Arequipa massif of coastal Peru:
Journal of the Geological Society, v. 136, p. 195–214 J. A. Salfity
Tankard, A. J., M. A. Uliana, H. J. Welsink, et al., 1995, Pje. N. Roldan 57
Tectonic controls of basin evolution in southwestern 4400 Salta
Gondwana, in A. J. Tankard, R. Suarez, and H. J. Welsink, Argentina
Tectonics and Stratigraphy of the Late Paleozoic
Paganzo Basin of Western Argentina and its
Regional Implications

Fernando Fernandez-Seveso Anthony J. Tankard


YPF SA Tankard Enterprises
Buenos Aires, Argentina Calgary, Alberta, Canada

Abstract

T he Carboniferous–Permian Paganzo succession straddles the Pampeanas, Precordillera, and Chilenia


terranes. Late Devonian–Early Carboniferous diastrophism of the Chañic event separated very different
early and late Paleozoic histories of basin formation. The Paganzo basin was initiated in the Visean by reactiva-
tion of old terrane boundaries. The early Paganzo consisted of a suite of discrete fault-controlled depocenters
interpreted as transtensional pull-apart basins linked to right-lateral displacement along major crustal faults.
Younger phases of basin formation were characterized by amalgamation of these various depocenters into a
single broad basin.
The Paganzo succession is divided into four supersequences by major hiatuses. These are the Guandacol,
Tupe, and lower and upper Patquía–De la Cuesta supersequences. Each is constructed by stacked unconfor-
mity-bounded depositional sequences. These four supersequences record the various stages of basin evolution.
The Guandacol sediments were deposited in isolated basins. Fieldwork shows a pattern of rapid subsidence
and stacking of coarse alluvial facies along basin-bounding faults. The characteristics of the finer grained strata
indicate a periglacial influence. The overlying Tupe supersequence suggests a gradual cessation of fault
activity as the various depocenters were yoked together. Tupe stratigraphy onlaps the Guandacol–Tupe
unconformity and buries some of the previous interbasin highs. In Patquía–De la Cuesta time, the Paganzo
basin had widened to its maximum extent. Significant transgressions are recorded in Tupe (Westphalian–
Stephanian) and Patquía–De la Cuesta (Artinskian and Kazanian) stratigraphy. Extensive geochemical studies
show that Patquía source rocks are oil prone. Although indications are that the Paganzo basin is prospective, it
remains largely untested.
Regional studies show that the strike-slip faults that controlled Carboniferous basin development in north-
western Argentina diverge northward where they become involved in the Chaco salient of the Bolivian Andes.
The Tupambi-Tarija and Escarpment sequences of Bolivia are broadly contemporaneous with the Guandacol
and Tupe stratigraphy of the Paganzo basin. They share similar depositional characteristics typical of rapidly
subsiding transtensional basins, including stacked alluvial facies, thick debris flow diamictites, massive soft
sediment deformation, and dewatering structures. The Escarpment Formation represents an expansion of the
earlier Tupambi-Tarija depocenters and contains an anastomosing drainage system.

Resumen

L a sucesión carbonífera-permica de Paganzo se dispone sobre los terrenos de Pampeanas, Precordillera y


Chilenia. La fase Chañica del diastrofismo Devónico tardío-Carbonífero temprano separó las singulares
y diferentes historias de formación de las cuencas del Paleozoico inferior y superior. La cuenca de Paganzo se
inició en el Viseano por reactivación de suturas de antiguos terrenos. El Paganzo temprano consistió de un
conjunto de discretos depocentros controlados por falla, los que se interpretan como cuencas “pull-apart” rela-
cionadas a desplazamientos laterales dextrógiros a lo largo de importantes fallas corticales. Las fases mas
jóvenes de la formación de la cuenca se caracterizaron por la amalgamación de varios depocentros en una sola
cuenca amplificada.
La sucesión estratigráfica de Paganzo es dividida en cuatro supersecuencias segun las principales discor-
dancias reconocidas. Estas fueron denominadas Guandocol, Tupe, y Patquia-De la Cuesta inferior y superior, y
caracterizan dose por el apilamiento estratigráfico de secuencias depositacionales acotadas por discordancias.
Estas cuatro supersecuencias reflejan los varios estadíos de evolución involucrados. Los sedimentos de la
supersecuencia Guandacol fueron depositados en cuencas aisladas, con tasa de subsidencia elevada y acumu-

Fernandez-Seveso, F., and A. J. Tankard, 1995, Tectonics and stratigraphy of the late 285
Paleozoic Paganzo basin of western Argentina and its regional implications, in A. J.
Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG
Memoir 62, p. 285–301.
286 Fernandez-Seveso and Tankard

lación de facies aluviales gravitacionales gruesas, a lo largo de márgenes controlados por falla. Los atributos
presentes en los estratos de granulometrías finas indican influencia periglacial. La suprayacente supersecuencia
Tupe sugiere un gradual cese de actividad en las fallas, a partir de que varios depocentros comenzaron a
unirse. Los estratos del Tupe traslapan sobre la discordancia que los separa de la supersecuencia Guandacol y
cubren algunos de los preexistentes altos intracuencales. Para el tiempo Patquía-De la Cuesta, la cuenca de
Paganzo ya había alcanzado su máxima extensión areal. Se registraron significativas transgresiones en los
estratos del Tupe (Westfaliano-Stefaniano) y Patquía-De la Cuesta (Artinskiano, Kazaniano). Numerosos
estudios geoquímicos señalan que la roca madre de Patquía es generadora de petróleo. Aunque la cuenca de
Paganzo presenta evidencias de ser prospectiva, la misma aun se encuentra poco investigada.
Estudios regionales muestran que las fallas de desplazamiento lateral que controlaron las cuencas del
Carbonífero en el noroeste Argentino, divergen hacia el norte donde se involucran en la saliente del Chaco de
los Andes bolivianos. Las secuencias de Tupambi-Tarija y Escarpment son, en sentido amplio, contemporáneas
con la estratigrafía de Guandacol y Tupe de la cuenca de Paganzo. Estas comparten similares características
depositacionales, típicas de cuencas transtensionales de rápida subsidencia, incluyendo acumulación de facies
aluviales, espesas diamictitas de flujos de detritos, deformación en masa de sedimentos, y estructuras de licue-
facción. La Formación Escarpment con un sistema de drenaje anastomosado, representa una expansión de los
preexistentes depocentros del conjunto Tupambi-Tarija.

INTRODUCTION
The Paganzo basin originated in Early Carboniferous
time as a result of intense tectonism that disrupted the
Precambrian granitic and metamorphic basement and its
lower Paleozoic cover (Gordillo and Lencinas, 1979;
Baldis et al., 1982; Caminos, 1985). The basement already
had a marked anisotropy attributed to terrane accretion
and Neoproterozoic–Cambrian diastrophism of the
Brasiliano event (Figure 1) (Ramos et al., 1986; Tankard et
al., 1995). The Paganzo basin of west-central Argentina
covered 145,000 km2 and spanned about 100 m.y. of the
stratigraphic record. About 4500 m of sediments accu-
mulated.
Ramos (1988a) has interpreted the basement of
southern South America as a complex mosaic of cratonic
blocks. The Paganzo basin straddles a large part of the
Pampean terrane in the east and the Precordillera and
Chilenia terranes in the west (Figures 1, 2). Predictably,
this basement framework resulted in a highly segmented
depositional landscape. The western depocenters,
consisting of the Calingasta-Uspallata and Río Blanco
basins (Amos, 1972), are separated from eastern
depocenters by the discontinuous proto-Precordillera
ridge (Amos and Rolleri, 1965; Baldis and Chebli, 1969).
The various depocenters were intermittently connected
across this ridge (Lopez-Gamundi et al., 1989; Milana et
al., 1987). This tectonic setting was linked to subduction
in the Carboniferous (Ramos et al., 1986; Mpodozis and
Ramos, 1989).
The predominantly terrigenous clastic fill of this basin
complex reflects successive episodes of deposition in a
variety of marine and continental environments. The
repeated stacking of these depositional sequences,
intraformational deformation, and the numerous uncon- Figure 1—Relationship between Precambrian–early
Paleozoic terranes (modified after Ramos, 1988a) and
formities of various scales suggest a stratigraphic Phanerozoic basin development. The Paganzo basin spans
response to the basin-forming tectonic processes and to three terranes where their boundaries converge. AA,
sea level fluctuations. Although there may be a structural Arequipa–Antofalla terrane; CH, Chilenia terrane; P,
justification for these depocenters or subbasins, the Pampean terrane; PC, Precordillera terrane; PA, Patagonia
overall continuity of the stratigraphic cover suggests that terrane; RP, Rio de la Plata terrane. White rectangle marks
these depocenters were yoked together for much of their study area.
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 287

Figure 2—Present structural framework of Precordillera


fold and thrust belt and Pampean basement, a tilt-block
basin province that has inverted the Carboniferous–
Permian Paganzo basin. Stippled areas are modern
depressions or valleys (bolsones); the curved lines trace
topographic relief and the principal inversion faults. The
broken line shows the extent of amalgamated Paganzo
basin complex. A–B shows regional cross section of Figure 3—Stratigraphic column and sequence framework
Figure 15. B, Bola Hill; BE, Bermejo lineament; HU, Huaco; for the Paganzo succession. Apices of triangles point in
ISCH, Ischigualasto basin; LS, Las Salinas; MA, Malimán; direction of overall sequence fining. Other symbols
MAS, Salinas de Mascasin; ML, Mogna–Las Salinas ridge; indicate the presence of marine fossils or plants.
MZ, Malanzán; P, Puerta de las Angosturas; PAS, Pasleam
syncline; PP, Protoprecordillera ridge; SA, Sañogasta; SE,
Santiago del Estero; VB, Veladero–Bola hills; VF, Sierra de
Valle Fértil.
This paper addresses the stratigraphic history of the
Paganzo basin during the Carboniferous–Permian, a
history. The easternmost Pampean terrane formed an time when Gondwana was subjected to one of the
area of high relief between the Paganzo and Chaco- world’s great ice ages. The approach involves a detailed
Paraná basins, shedding sediments into both (Alvarez et sequence stratigraphic analysis as a guide to the tectonic
al., 1990). evolution of the basin. The intention is to expand on
The Paganzo succession consists of large-scale uncon- earlier studies (e.g., Scalabrini-Ortiz, 1973; Vásquez et al.,
formity-bounded supersequences that express the broad- 1981; Salfity and Gorustovich, 1983; Archangelsky, 1986;
scale tectonic evolution of the basin and a stacked suite of Lopez-Gamundi et al., 1989; Fernandez-Seveso et al.,
smaller scale depositional systems or sequences 1988, 1990, 1993; Perez et al., 1993). Fernandez-Seveso et
(Fernandez-Seveso et al., 1990, 1993). The unconformity- al. (1993) provide a more detailed account of the
bounded supersequences form the Guandacol, Tupe, sequence classification and their correlation. The
Patquía, and De la Cuesta formations (Figure 3). Carboniferous Paganzo stratigraphy is compared to its
Mapping of these supersequences is based on surface counterparts in the Eastern Cordillera and sub-Andean
stratigraphic correlation and seismic interpretation. foothills of southern Bolivia.
288 Fernandez-Seveso and Tankard

basis, the Carboniferous–Permian stratigraphy in the


Paganzo basin consists of 15 depositional sequences
grouped into four supersequences by regional unconfor-
mities (Figure 3). Because of the nature of tectonism and
basin subsidence, these sequences and supersequences
are not uniform. The supersequences have a 12–34 m.y.
periodicity and thicknesses that range from 600 to
2000 m. In contrast, the depositional sequences have a
periodicity of 3–10 m.y. and a 100–600 m thickness range.
At the smallest scale of investigation, the basic genetic
units (lesser order sequences and parasequences) vary
up to 3 m.y. in duration and 250 m in thickness (Fer-
nandez-Seveso et al., 1993). This Paganzo stratigraphy is
reflected in subsidence curves.

STRATIGRAPHIC SETTING
The overall tectonostratigraphic reconstruction of the
Paganzo basin (Figure 4A) is derived from stratigraphic
relationships, facies characteristics, and seismic interpre-
tation (Figures 5, 6). The Visean-Namurian Guandacol
supersequence (G1–G4 of Figure 4A) is characterized by
Figure 4—Reconstruction of Paganzo basin evolution. (A) an asymmetric basin fill in which coarse facies are
Stratigraphic relationships based on stratigraphic sections,
stacked along the margins of the basins. Subsidence was
field mapping, and seismic interpretation (Figures 5, 6). (1)
The Visean–Namurian Paganzo basin consisted of discrete initially rapid enough to allow stacking of coarse clastics
fault-controlled depocenters. Coarse facies of the along the principal basin-forming faults of an extensional
Guandacol supersequence are stacked vertically adjacent or transtensional system (subsidence rate greater than
to basement-involved faults. (2) The Westphalian–Asselian sedimentation rate). In contrast, the overlying West-
Tupe supersequence records progressive amalgamation of phalian-Asselian Tupe supersequence (T1–T4 of Figure
these isolated basins to form a single basin complex with 4A) had an onlapping relationship that was more wide-
an irregular floor. (3) The late Asselian–Tatarian spread, locally blanketing the interbasin highs for the
Patquía–De la Cuesta supersequences were deposited in a first time.
uniformly subsiding intracontinental sag when the basin Each sequence is characteristically upward fining
was widest. (B) Detail of idealized Guandacol sequence
from gravel-rich and conglomeratic facies to thinly
deposited in a pull-apart basin. Each sequence is charac-
terized by stacking of coarse facies against an active laminated shales, indicating a repetition of similar depo-
basin-bounding fault. Distal facies include varved lacus- sitional processes (Figure 4B). Each has a fault-controlled
trine shales with evidence of turbiditic underflows and stratigraphy.
turbidites. (Modified after Fernandez-Seveso et al., 1993.)
Guandacol Supersequence
METHODOLOGY The Guandacol supersequence contains an Early
Carboniferous flora (Andreis and Arrondo, 1974). It is up
The study area includes the greater part of the basin to 1825 m thick in outcrop. The Guandacol superse-
outlined by Salfity and Gorustovich (1983) between quence was deposited beginning in the late Visean as a
about 27° and 31° S lat (Figure 2), augmented by a few suite of rapidly subsiding depocenters in which sedi-
control points in the Rio Blanco and Calingasta-Uspallata mentation was fault controlled. Subsidence was intermit-
basins. The data base includes 44 detailed stratigraphic tent and at times rapid. This is reflected in vertical
sections measured in outcrop, which were subsequently stacking of coarse facies adjacent to a basin-bounding
integrated with interpreted reflection seismic profiles. A fault, debris flows, and massive synsedimentary defor-
conventional sequence stratigraphic analysis was mation structures. The rate of subsidence exceeded the
applied. The aim was to establish the pattern of basin rate of sediment supply. At other times, the basin was
subsidence on the basis of variations in the nature and largely starved of coarser detritus, and more argillaceous
distribution of depositional systems and sedimentary lacustrine material draped the entire floor of the basin.
facies. The ages of the sequences and their bounding Rhythmic lacustrine deposits, faceted clasts, and giant
surfaces are based on flora, palynomorphs, microfossils, dropstones suggest a periglacial environment during the
and larger marine invertebrates. late Visean and Namurian.
Our interpretation of the marine and terrestrial stratig- The Guandacol supersequence is divided into four
raphy of the Paganzo basin emphasizes relative changes depositional sequences by bounding erosional unconfor-
of base level that are believed to encompass the effects mities (Figures 3, 4A, 7). Each sequence fines upward
both of relative sea level and basin subsidence. On this from gravel-rich and conglomeratic facies to thinly
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 289

Figure 5—Unmigrated seismic line 9052 from Bermejo valley showing erosional truncation of pre-Carboniferous basement
(reflector 1), the Paganzo–Ischigualasto basin complex (between reflectors 1 and 2), and a zone of high-amplitude contin-
uous reflectors attributed to the Cretaceous–Paleogene section (above 2). Reflector 3 separates the Triassic Ischigualasto
basin from the Paganzo basin. The thick section above 4.5 sec on the eastern side is Neogene fill in a tilt-block foreland
basin. See Figure 2 for location.

Figure 6—North-south oriented line 9053 crossing previous line 9052 in Figure 5. Note scale change. See Figure 2 for location.

laminated shales, indicating a repetition of similar depo- debris flows and mudflows.
sitional processes (Figure 4B). Each has a fault-controlled • Laterally continuous sandstones with sharp bases
stratigraphy. The depositional facies (Figures 4B, 7) separated by thinly interbedded mudstones and
include the following: sandstones (Figure 8), locally with massive
slumping. Fernandez-Seveso et al. (1993) have
• Gravel-rich, conglomeratic sandstones with attributed these sediments to turbiditic underflow
multiple channel scours proximally. These have processes.
been interpreted as fan deltas with associated • Laminated shales and siltstones with sandstone
290 Fernandez-Seveso and Tankard

Figure 7—Guandacol supersequence consisting of four


stacked sequences. Symbols along stratigraphic column
indicate glaciogenic intervals (dropstones and varvelike
rhythmites). Also shown are TOC and C15+ extractable
bitumen suggesting low to moderate potential gas-prone
sediments.

interbeds and slump structures. The finer grained


lithologies are markedly varved and contain drop-
stones, some of them very large (Figure 8). Plant Figure 8—Lacustrine mudstones and siltstones deposited
remains, palynomorphs, and Lioestherias occur; as varved couplets with thicker turbidite interbeds and
there is no evidence of marine paleoenvironments large granitic dropstone. Dropstone is about 1 m in
in the Bermejo and Pasleam depocenters (Figure 2). diameter. Guandacol supersequence.
These rocks are believed to have been deposited in
lacustrine environments that were locally influ-
enced by turbiditic flows.

These stacked fining-upward Guandacol sequences


have a lateral continuity on outcrop scale (Figure 9), but
regionally they have a limited distribution. The most
widespread unit is the lacustrine shale–siltstone facies
tract that draped and interfingered with the slumped
mudflows and turbiditic sandstones (Figure 4B).
Renewed sediment influx was distributed by sub-
aqueous gravity flows that formed channelized and
nonchannelized depositional lobes typical of classic
turbidite systems (see Mutti, 1985).

Tupe Supersequence
The Tupe supersequence is an Upper Carboniferous
(Westphalian) to Lower Permian (Asselian) succession
separated from the Guandacol by a major unconformity
(Figure 3). Like the Guandacol, the Tupe succession Figure 9—Guandacol–Tupe succession at Bola Hill. Lower
consists of four sequences separated by unconformities. part is fining-upward Guandacol sequences. Lower
Each sequence is a progradational complex that reflects, Patquía–De la Cuesta supersequence at crest is overlain
to varying degrees, deposition in fluvial, lacustrine, and by Triassic basalts. See Figure 2 for location.
marginal marine environments (Figure 10). The Tupe
succession is as thick as 1285 m. Permian age (A. Chaia, 1993, personal communication),
The age of the Tupe succession is believed to be whereas the invertebrates in Tupe sequence 3 have
predominantly Late Carboniferous (Westphalian– Stephanian affinities (N. Sabattini, 1993, personal
Stephanian) on the basis of palynomorphs and fossil communication).
plants (O. Arrondo, D. Ganuza, G. Leunda, and E. Morel, Tupe sedimentation initially reoccupied old depocen-
1993, personal communications). The ostracod genus ters established in Guandacol time. However, the Tupe
Bairdiacypris indicates a Late Carboniferous–Early stage of basin subsidence differed from the earlier
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 291

carbonate shorezone deposits with diverse marine inver-


tebrate assemblages. A variety of bioclastic and
biohermal facies of low energy littoral origin overlie
sequence 2 and Precordillera basement. Ash tuffs occur
locally.
Sequence 4 reflects progradation of the shoreline as
the basin filled with braidplain deposits. Rare dropstones
and rhythmites occur in some lacustrine deposits
(Rodriguez and Fernandez-Seveso, 1988). Marginal
deltaic deposits locally form laterally continuous
sandstone sheets and channel fills. Fossils include
ostracods and casts of invertebrates. The upper part of
the sequence is characterized by subtle truncations and
toplap relationships.

Figure 10—Tupe supersequence consisting of four stacked


sequences. Also shown are TOC and C15+ extractable
Lower Patquía–De la Cuesta
bitumen indicating lean gas-prone sediments, not prospec- Supersequence
tive source rocks. Symbols along stratigraphic column
The lower Patquía–De la Cuesta (LPD) is a classic
indicate glacigenic intervals (dropstones and varvelike
rhythmites) and limestones. Lower Permian red bed succession characterized by
alluvial fan, fluvial, and playa lake lithofacies that
encroached on a shallow marine basin (Figures 3, 16)
Guandacol stage in two important ways. First, sedimen- (Fernandez-Seveso et al., 1993). These sediments are
tation overlapped the margins of the old depocenters, as contained within three stacked sequences that reflect an
well as the interbasin highs (Figures 4A, 11), suggesting overall transgressive-regressive history. The upper part
that by Westphalian time those depocenters were of the LPD consists of thick eolianite sandstones
gradually yoked together in a regionally subsiding basin (Limarino and Spalletti, 1986). The lower Patquía–De la
(Figures 4A, 6). Second, the basin was subjected to Cuesta succession is up to 620 m thick.
periodic inundation by the sea, as reflected in the trans- A late Asselian–Kungurian age is inferred from the
gressive-regressive stratigraphy and the marine biota. magmatic flows and pyroclastics, which are attributed to
The constituent depositional sequences were irregu- the Choiyoi magmatic episode (Caballé, 1990). The
larly distributed over ridges and interbasin highs. The volcanism occurred mainly toward the west on the
proto-Precordillera, the Mogna–Las Salinas ridge (Perez Chilenia terrane. Early volcanism was mainly andesitic,
et al., 1993), and the Valle Fertil and Sañogasta margins but changed upward to rhyolitic composition. Palyno-
(Figure 2) remained active. In the northeast and east, logic assemblages (LPD2) (Figure 13) suggest a broad
overlapping relationships suggest that new, albeit Late Carboniferous–Early Permian age range (G.
shallow, depositories had formed (Figure 4A). The basin Leunda, 1993, personal communication).
expanded its margins farthest during middle Tupe time Sequence 1 onlaps a low relief regional unconformity
(sequences T2 and T3). Fluvial and lacustrine deposi- above the previous Tupe succession (Figure 4A). Conti-
tional systems dominated the eastern part of the basin nental sedimentation was widespread and displaced the
over Pampean basement. The succession thickened Tupe sea, even as far as the western edge of the Pre-
westward where fluvial and sandy deltaic systems cordillera terrane. This sequence is dominated by chan-
interfingered with marginal marine deposits. nelized conglomerates and sheet sandstones interpreted
Sequence 1 reoccupied previous Guandacol depocen- as distal alluvial fan and braided fluvial deposits.
ters. Sediments were distributed by subaqueous channel Whereas the eastern parts of the basin contain stacked,
and turbidite processes in front of sandy fan delta coarse, braided fluvial deposits, the northern area is char-
systems and were locally subjected to contemporaneous acterized by widespread mudstone and interbedded
slumping on a massive scale (Figure 12). sandstone deposits cut by lenticular channel sandstones.
Sequence 2 overlies the Precordillera terrane (Figure We infer a low relief landscape of playa lakes and
11) (Perez et al., 1993) and the western margin of the ephemeral streams. Volcaniclastic material and basalt
Pampean terrane. It consists of three overlapping flows are observed locally (Azcuy and Morelli, 1970).
packages that were built basinward by braided fluvial Sequence 2 reflects the paleogeography of a marine
and Gilbert-type fan delta systems. Basinal and interdis- transgression. Over a large part of the Precordillera
tributary bay mudstones are varved and contain drop- platform in the north, shore zone wave and tidal influ-
stones and faceted clasts. ences were important, depositing relatively mature
Sequence 3 records the maximum expansion of the reworked sandstones. This highstand systems tract
Tupe paleogeography (Figure 4A). This sequence onlaps contains Tasmanaceas and Acritarcos microplankton (e.g.,
and oversteps interbasin highs. A depositional history Leiosphaeridium sp.) and algae, indicating restricted
similar to the previous sequence prevailed, including brackish environments. In contrast, the eastern edge of
overlapping depositional lobes and local intraforma- the basin above the Pampean basement was buried by
tional disconformities. There are also terrigenous and braided fluvial and argillaceous floodplain deposits.
292 Fernandez-Seveso and Tankard

Figure 11—Tupe succession (sequence 2) onlapping interbasin high of Precordillera basement, Veladero Hill. White sand-
stones are fluvial. See Figure 2 for location.

Sequence 3 varies up to 230 m thick and consists of alluvial fan, debris flow, and braided sheetflood processes
well-sorted eolian sandstones with intraformational that suggest a period of slightly higher rates of subsi-
shale facies of possible interdune origin (Figure 13). dence. In the inner part of the basin, these alluvial facies
are associated with the argillaceous deposits of playa and
Upper Patquía–De la Cuesta perennial lakes. These lakes were probably the result of
Supersequence impounding by alluvial systems or erosional deflation
over the axis of the basin. Restricted marine incursions are
The upper Patquía-De la Cuesta (UPD) is a succession also inferred from microfossils (e.g., acritarchs and
of terrigenous clastic deposits that formed in ephemeral ostracods) and limestone interbeds. The lacustrine to
river, playa lake, perennial lake, and marginal marine restricted marine deposits contain oil-prone bituminous
settings (Figure 13) (Fernandez-Seveso et al., 1993). This shales (Figures 13, 14). We infer a restricted marine and
is an Upper Permian succession consisting of four fringing coastal plain paleogeography with a warm,
stacked depositional sequences. The stratigraphy is most humid climate. This lower sequence is equivalent to the
complete in the northern part of the basin. The UPD Vitiacua Formation of southern Bolivia (Sempere, 1995).
terminates with a major unconformity that truncated as The margins of the second depositional sequence
deeply as the Tupe supersequence in the central part of were characterized by small fluvial channels that fed
the basin where the Triassic Ischigualasto basin subse- shoalwater mouth bars. Finer grained interdistributary
quently developed (Figures 5, 9), suggesting that a bay and marsh deposits were widespread and contain
period of structural inversion separated the Permian and plant remains, palynomorphs, and terrestrial inverte-
Triassic basins (Perez et al., 1993). The upper Patquía is brate fossils such as Anthracosiasea sp. (G. Leunda, 1993,
up to 730 m thick. personal communication).
The age of the upper Patquía is believed to be early A regional unconformity forms the base of the third
Late Permian on the basis of palynomorphs in the lower depositional sequence. It contains eolianites, fluvial
two sequences (Aceñolaza and Vergel, 1987; G. Leunda, deposits, and shallow lacustrine deposits with
1993, personal communication). interbedded sheetflood sandstones (Figure 13). Sequence
The lowest depositional sequence was deposited 4 at the top of the Paganzo succession is dominated by
unconformably on a lower Patquía eolianite by distal sandstones attributed to ephemeral river deposition.
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 293

Figure 13—Lower and upper Patquía–De la Cuesta super-


sequences comprising a total of seven sequences. VV,
Figure 12—Turbiditic sandstones with a massive slump Basalt flows. Also shown are TOC and C15+ extractable
structure, Tupe sequence 1, Bola Hill. Slump is about 5 m bitumen suggesting oil-prone source rocks with high
thick; see person at right for scale. hydrogen index.

INTERPRETATION
The Paganzo basin fill reflects a complex tectonic and
paleogeographic setting involving intermittent fault-
controlled reactivation of old basement terranes and
extrabasinal sea level and climatic influences (Tankard et
al., 1995).
Much of the Carboniferous Guandacol and Tupe
deposition occurred in periglacial settings as evidenced
by the varved lacustrine shales with dropstones, faceted
clasts, diamictites interpreted as muddy debris flows, the
predominance of humid braided fluvial deposits, and the
plant ecology (O. Arrondo, E. Morel and E. Ganuza,
1993, personal communication). There is little direct
evidence for glaciation in the Paganzo basin. However,
glacial pavements and tillites have been recorded in the Figure 14—Potential oil-prone source rocks (two dark
proto-Precordillera and Mogna–Las Salinas ridges (Figure bands) in upper Patquía (sequence 1). Sierra de Narvaez
2) (Lopez-Gamundi et al., 1989; Milana and Bercowski, ridge is in the background. Puerta de las Angosturas
1990) where there was topographic control. In contrast, section, northern part of Paganzo basin. See Figure 2 for
post-Tupe Late Carboniferous–Early Permian coastal location.
elevations are believed to have had a cold, temperate
paleoclimate, as suggested by foraminifera and mollusc
fossils in the Huentelauquén Formation on the coast of
central Chile (Rivano and Sepúlveda, 1985). By lower
Patquía time, climatic warming had set in. Glacial envi- coincided with accumulation of bituminous black shales
ronments are not characteristic of other Early Permian in the Paganzo basin (Figure 14) and the Vitiacua
Gondwana basins (Zalan et al., 1987; Visser, 1991; Eyles, Formation of the Tarija basin. Organic-rich shales of this
1993). However, Zalan et al. (1987) and Russo et al. (1986) age also occur in the Iratí Formation of the Paraná basin
have documented Early Permian glaciation in the Paraná and the Whitehill Formation of the Karoo basin (Franca
and Chaco-Paraná basins, respectively. et al., 1995; Tankard et al., 1995, their figure 23).
By the end of early Patquía time, the climate was In Gondwana reconstructions, the Late Carbonif-
universally arid, and desert dune fields were wide- erous–Early Permian Paganzo basin was located
spread, as was the case for the fluvial-eolian Cangapi between 30° and 45° S lat, compared to 70° S lat in the
Formation in the Tarija basin of northwestern Argentina. Early Carboniferous (Smith et al., 1981; Scotese and
This warming trend also resulted in a pronounced sea Barrett, 1990). Consequently, glacial paleoclimates over-
level rise during the late Kungurian or Kazanian printed Guandacol deposition; these were the late
(Sempere et al., 1992; Sempere, 1995). This episode Visean–early Namurian and late Namurian glacial peaks
294 Fernandez-Seveso and Tankard

Figure 15—Tectonic model for evolution of Paganzo basin complex. Initially the Paganzo developed as a number of discrete
basins by reactivating old basement fabrics. In the Late Carboniferous (Tupe time), these were joined together in a single
wide basin. (Modified after Fernandez-Seveso et al., 1993.)

(J. C. Crowell, 1990, personal communication). Eyles basins were predominantly terrestrial depocenters. The
(1993) has reviewed this geology and attributes this early Lower Carboniferous stratigraphy in the Paganzo basin
glaciation to oblique collision, rotation of large crustal is poorly understood, but lithostratigraphic, biostrati-
blocks, and glaciation on tectonically uplifted areas. graphic, and radiometric age data place it above the
During the Late Carboniferous, the world became Chañic unconformity (Polansky, 1970; Ramos and
distinctly colder as the Gondwana ice sheets expanded Ramos, 1979). In the Precordillera, Carboniferous strata
(Dickins, 1993). However, the characteristics of sedimen- overlie folded pre-Carboniferous rocks.
tary facies show that the upper Tupe Paganzo basin was Guandacol deposition occurred in small, isolated
already at warmer latitudes. These latitudinal differences depocenters (Figure 15). Despite their small size, a rela-
also explain why Carboniferous sedimentation rates tively thick succession (nearly 2000 m) of sandstones and
were higher in the Paganzo and Tarija basins than in the conglomerates were stacked adjacent to a basin-
Paraná and Chaco-Paraná basins (see Russo et al., 1986; bounding fault (Figure 4), suggesting that the rate of
Zalan et al., 1987); the former basins were periglacial at subsidence exceeded the rate of sediment supply.
that time. Massive synsedimentary deformation (Figure 4)
Much of Paganzo deposition occurred in a succession supports this interpretation. Braided alluvial fan and
of transgressive-regressive cycles. An abrupt lowering of debris flow deposits suggest a rugged relief. The
sea level from the late Namurian to early Westphalian, Guandacol depocenters strongly resemble the transten-
together with a change in the rate of subsidence, resulted sional pull-apart basins of strike-slip zones (e.g., Nilsen
in the unconformity between the Guandacol and Tupe and McLaughlin, 1985). Ramos (1988b) has interpreted
successions. Maximum flooding occurred during the strike-slip faults in Mendoza Province in a right-lateral
Westphalian and Stephanian (Tupe sequences 2 and 3). sense due to displacement of the Chilenia terrane. Thick
There was a smaller advance in the Artinskian (lower sedimentary sections above the Pampean-Precordillera
Patquía) (see Figure 16). The Upper Permian (Kazanian) and Precordillera-Chilenia sutures contain Visean fossils,
organic-rich or bituminous shales in the upper Patquia, thus dating the initiation of Guandacol basin formation
Vitiacua, and Irati formations all indicate a warm, post- to the Visean.
glacial climate and pronounced transgression (Sempere Fault-controlled subsidence initially formed small
et al., 1992). half-grabens in the eastern part of the Guandacol basin.
A long history of early–middle Paleozoic basin These pull-apart basins subsided intermittently in a
evolution ended with the Late Devonian–Early Carbonif- series of rapid spurts over a 35-m.y. interval. This episode
erous Chañic orogeny (Aceñolaza and Toselli, 1976; was succeeded by the widespread Tupe complex as the
Salfity and Gorustovich, 1983). This diastrophic phase is small basins of the transtensional system ceased activity
expressed throughout southwestern Gondwana by a and the various depocenters were amalgamated in a
substantial hiatus. The basin deposits on either side of uniformly subsiding depression. This change in basin
this regional unconformity are very different. The early dynamics is marked by a regional unconformity. The
Paleozoic basins were largely underfilled marine basins. Tupe succession onlapped this unconformity and over-
In contrast, the Carboniferous and Permian successor lapped the older interbasin highs and, for the first time,
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 295

Figure 16—Back-stripped tectonic and basement subsi-


dence curves for Permian–Carboniferous Paganzo basin
showing three stages of subsidence: rapid transtensional
subsidence, regional subsidence, and increased rates of
subsidence in the Permian attributed to stress relaxation.
The geohistory plot includes outcrop data, petrophysical
logs, and seismic data. (Modified after Fernandez-Seveso
et al., 1993.)

expanded the Paganzo basin across the Pampean terrane


(Figure 15).
A third style of subsidence was established in early
Patquía–De la Cuesta time. Patquía sediments record Figure 17—Interpretation of regional tectonic setting of
maximum expansion of the basin as they onlapped even Paganzo basin complex. The structural framework that
farther across the basement. There was a general lack of formed the broadscale basin architecture probably
resulted from right-lateral displacement between terranes
relief in the basin, although some sedimentary facies
and northeastward-splaying of the Sierra de Valle Fértil
suggest an increase in the rate of subsidence. This phase fault in the passive Pampean terrane. Rhombic pull-apart
most closely resembles a shallow epeiric basin or intra- basins were probably associated with this diverging strike-
continental sag. This pattern persisted until inversion at slip system.
the end of the Permian marked a tectonic reorganization
and the onslaught of Triassic extension by orogenic
collapse (Tankard et al., 1995). because of a smaller wrench component. This extensional
This tectonostratigraphic interpretation is summa- stage culminated in extrusive magmatism along the
rized in back-stripped tectonic and basement subsidence terrane boundaries in early Patquia time. Slightly higher
curves (Figure 16) (see Bond and Kominz, 1984). The subsidence rates in the Late Permian may be attributed
stratigraphy has been decompacted taking into account a to extensional relaxation (see Bally and Snelson, 1980), as
7000-m post-Permian section and empirical porosity to Mpodozis and Kay (1990) and Llambias and Sato (1990)
depth relationships. The greatest inaccuracy in the have suggested for the Gondwana margin. Upper
Paganzo subsidence curves is their imprecise chronology. Patquía alluvial fan and debris flow facies support this
The back-stripped subsidence curves record three interpretation.
styles of subsidence (Figure 16). The first phase of subsi- The Carboniferous Paganzo basin straddled the
dence, spanning about 30 m.y., was the most rapid and Pampean, Precordillera, and Chilenia terranes (Figure
resembles an open rift style (Meyer, 1991; Williams, 17). The basin initially took advantage of terrane bound-
1995). Field and seismic mapping suggest that aries and the relative strengths of the terranes them-
Guandacol depocenters subsided as transtensional pull- selves. The Guandacol depocenters developed as
apart basins related to a dextral strike-slip system. In the transtensional pull-apart basins. Ramos (1988b) docu-
Paganzo basin, the conglomeratic-volcanic complex of ments a dextral shearing system. Major displacements
the Punta del Agua Formation (Gonzalez and Bossi, occurred along the western boundaries of the Pampean
1986) is the only evidence of magmatism at this stage. and Precordillera terranes. The Sierra de Valle Fértil fault
The small amount of magmatism during the dynamic system is a major crustal structure and terrane boundary
stage of subsidence of an intracratonic basin facing a that was reactivated in the Paleozoic (Baldis et al., 1982).
convergent plate boundary has been noted elsewhere We believe that horizontal displacement was not
(Mann et al., 1983; Christie-Blick and Biddle, 1985). The restricted to the terrane boundaries, but that secondary
succeeding Tupe part of the curve is flatter, perhaps strike-slip faults and associated normal faults related to a
296 Fernandez-Seveso and Tankard

shear couple dissected the more passive Pampean ppm, high 4000 ppm). The organic matter is
basement (Figure 17). composed mainly of agglutinated amorphous
This system of composite wrench faults and a types with a low woody component. Some of the
network of transtensional and transpressional compo- best samples contain botryococcus algae and disin-
nents is common in areas of oblique convergence (Ryan tegrated pollen grains and occasionally have more
and Coleman, 1992). These composite plate boundaries than 85% spongy amorphous material. Microscopic
may also result in rotation of crustal blocks. The Pie de and chromatographic analyses indicate lacustrine
Palo block within the Precordillera terrane (Figure 2) type I and II kerogens with sapropelic affinities.
may be an example. Rapalini (1989) has already docu- These geochemical studies suggest good to very
mented oblique convergence and block rotation along good potential for oil generation.
this margin of Gondwana. Subduction during the
Carboniferous has been described by Ramos et al. (1986) The upper Patquía–De la Cuesta shales were
and Mpodozis and Ramos (1989). Fielding and Jordan deposited in lacustrine and restricted marine environ-
(1988) have addressed the Recent deformation of the ments during a prominent transgressive phase that was
Precordillera and Sierras Pampeanas in the Andean influenced by a warm, temperate postglacial climate (see
foreland as a tilt-block province. Dickins, 1993). The highest source potential occurs in
shales near the base of the upper Patquía–De la Cuesta
(Figure 14). These oil-prone shales are particularly wide-
spread in the northern part of the basin; this facies is
HYDROCARBON POTENTIAL OF more arenaceous toward the south. Generated oil is
PAGANZO BASIN observed in outcrop where it fills pores, cavities, and
fissures and is even associated with andesite sills.
Routine geochemical studies, including total organic The results of fieldwork, exploration wells, and
carbon (TOC), C15+ extractable bitumen, optical identifi- seismic interpretation suggest that Upper Permian
cation of organic macerals, elemental analyses, gas chro- source rocks are probably widespread in Paganzo strati-
matography, thermal alteration indices (TAI), and graphic units that are now preserved in Tertiary tilt-
vitrinite reflectance have been performed on hundreds of block basins. There is also source rock potential in the
surface samples gathered from 32 sections, as well as on Triassic Ischigualasto basin (Figure 5) similar to the Cuyo
cuttings samples from the Salinas de Mascasín well basin. Surface oil occurrences in the Paganzo basin have
(Figure 2) (Fernandez-Seveso et al., 1991). All argilla- been known for many years (Fernandez-Seveso et al.,
ceous units in the Paganzo basin were sampled and 1991; G. Kelly, 1993, personal communication). These
potential source rocks identified. The TOC and include asphalts and impregnated oil in Tertiary
extractable bitumen histograms (Figures 7, 10, 13) outcrops of Niquivil Viejo (Figure 2). Heavy oil has also
represent basinwide averages. Weathering alteration of been reported in wells in the Bermejo, Niquivil (northern
surface samples was adjusted as follows: TOC × 1.3 and part of Mogna–Las Salinas ridge), and Las Salinas areas.
C15+ extractable bitumen × 2. The following results were Except for wells drilled near the basin edge in the Salinas
obtained: de Mascasín, which reached the Tupe, no wells in either
the Paganzo or Ischigualasto basin have reached source
• Guandacol dark shales (Figure 7) have low to rock depth. All of the wells drilled in Tertiary sediments
moderate TOC values (range 0.1–2.5%, max. 5%). and continental red beds are of probable Cretaceous age.
The C15+ extractable bitumen values are low The Paganzo basin is still largely untested, although
(0–700 ppm). Woody and cuticular characteristics there is substantial evidence to suggest that it is
suggest a gas-prone succession. prospective.
• Tupe carbonaceous shales are not encouraging
(Figure 10). Sequences 2 and 4 deposited in lacus-
trine environments should have the best potential,
but TOC values are low to moderate (range REGIONAL PALEOGEOGRAPHIC
0.5–2.5%, max. 7%). The C15+ extractable bitumen is IMPLICATIONS
very low (0–200 ppm, locally 600 ppm). Organic
material consists of gas-prone woody structures, The sedimentary facies, stacking arrangements, and
herbaceous material, cuticles, and sporomorphs. distribution of the Paganzo Group reflect the geody-
Even marine mudstone facies are lean (see Villar namic setting of a suite of pull-apart basins. These basins
and Lopez-Gamundi, 1993). evolved after the basal Carboniferous Chañic dias-
• The lower Patquía–De la Cuesta mudstones were trophism by transtensional processes, perhaps related to
deposited in strongly oxic conditions. Organic dextral convergence of Chilenia (Ramos, 1988b). Indeed,
material includes microplankton, botryoidal algae, Lower Carboniferous Guandacol counterparts are well
sporomorphs, and woody structures. exposed in Paganzo latitudes (Archangelsky, 1986;
• The upper Patquía–De la Cuesta shows a dramatic Gonzalez and Bossi, 1986; Lopez-Gamundi et al., 1989;
increase in source potential (Figure 13), with good Cingolani et al., 1990).
TOC values (average 1.2%, max. 5%) and excellent The Paganzo basin is at one end of the Precordillera
C15+ extractable bitumen content (range 250–2000 thrust belt and tilt-block province that dominates north-
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 297

sediments, large-scale synsedimentary deformation, and


prominent paleovalleys (Starck et al., 1992). Reyes (1972)
has attributed this Machareti Group sedimentation to
fault-controlled subsidence in small rift basins.
Carboniferous rocks are widespread in the Andean
ranges and Chaco basin of southern Bolivia (Figure 18)
where they form prolific hydrocarbon reservoirs. The
succession in the Eastern Cordillera and sub-Andean
foothills is remarkably similar to that of the Paganzo
basin, except that the Bolivian examples have been
detached and transported piggy-back on thin-skinned
thrust sheets of Tertiary age. The Machareti and
Mandiyuti groups are stratigraphic equivalents of the
Guandacol and Tupe supersequences, respectively
(Figure 19).
In outcrop, the Saipuru, Tupambi, and Tarija forma-
tions (Machareti Group) are counterparts of the
Guandacol supersequence of northwestern Argentina,
but without the same amount of repetitive stacking of
depositional sequences (Figure 19). Their deposition is
ascribed to lacustrine, braided alluvial fan, and bayhead
delta settings.

• Lacustrine facies occur in each of the formations.


They are typically argillaceous and contain thin
turbiditic layers, rhythmites, and sporadic drop-
stones. Bioturbation and rooting occur locally. The
Saipuru contains large slump structures.
• Alluvial fan and braided fluvial facies dominate the
Tupambi Formation and also occur in the Tarija.
The coarse-grained and conglomeratic sandstone
Figure 18—Carboniferous paleogeographic reconstruction intervals consist of multiply stacked channel units.
showing tectonic framework (Mon and Salfity, 1995), Early Inclined depositional surfaces and cross beds are
Carboniferous depocenters of Bolivia (after Reyes, 1972),
and Mandiyuti paleovalleys after Tankard et al. (1995).
common. The Tupambi is characterized by thick
B, Bermejo; IA, Izozog Arch; M, Mendoza; S, Sucre; SC, sequences of debris flow diamictites, ubiquitous
Santa Cruz; SE, Santiago del Estero; T, Tucuman; V, Villa- ripple drift and dewatering structures, and massive
montes; and Z, Zudañez. soft sediment deformation. Together these suggest
accumulation in a rapidly subsiding or collapsing
depocenter. Furthermore, cross-bed azimuths at
western Argentina as far as the Bolivian border (Figure Zudañez are eastward directed, and those in Piray
18), after which it is subordinated by the eastward River outcrops are oriented toward the west.
surging thrust belt of southern Bolivia. Basement struc- • The upper part of the Tarija Formation contains
tures have controlled basin development in north- coarsening-upward progradational sequences that
western Argentina since the late Precambrian (Salfity et are locally associated with bioturbation and rooted
al., 1975; Willner et al., 1987; Mon and Salfity, 1995). surfaces. A lacustrine-restricted marine bayhead
These faults were repeatedly reactivated in the early delta setting is inferred. Marine fossils diagnostic of
Paleozoic and culminated in a giant transpressional pop- the Levipustula levis zone indicate the Visean–West-
up structure attributed to the Chañic orogeny. This Puna phalian transition (Trujillo Ikeda, 1989; Sempere,
arch is astride west- and east-vergent basement-seated 1995).
structures that flank sedimentary basins on either side of
the arch (Mon and Salfity, 1995, their figure 6). The Upper Carboniferous Escarpment Formation is
The post-Chañic sedimentary basins subsided on the Tupe counterpart. It is a cliff-forming unit consisting
either side of the Puna arch during the Somuncura phase of stacked fluvial sandstone complexes 400–500 m thick.
of extension (Mon and Salfity, 1995). Compared with the Unlike the Tupambi–Tarija sequence, the Escarpment
2000-m thickness of the Paganzo succession, the deposits have suffered far less synsedimentary deforma-
Carboniferous rocks of the northwestern Argentinian tion. In the sub-Andean foothills, the Escarpment sand-
Tarija basin are closer to 1000 m in thickness (Figure 19) stones are geographically associated with the older
(Mingramm et al., 1979). The Lower Carboniferous Tupambi rocks (M. Cirbian, 1992, personal communica-
Machareti supersequence is stratigraphically equivalent tion).
to the Guandacol. Like the Guandacol, the Tupambi and In the subsurface of the Chaco basin, exploration
Tarija formations are characterized by rapid dumping of seismic and well data show that Carboniferous
298 Fernandez-Seveso and Tankard

Figure 19—Regional stratigraphic correlations. Villamontes


section from Sempere (1995). Formation names are as
follows: Ca, Cangapi; El, Elvira; Es, Escarpment; Ic, Ichoa;
Iq, Iquiri; LM, Los Monos; LP, Las Peñas; Sa, Saipuru; ST,
San Telmo; Ta, Tarija; Tu, Tupambi; and Vi, Vitiacua.

sediments form a widespread cover. An anastomosing linked pull-apart basins. Guandacol deposition reflects
network of Escarpment paleovalleys has been mapped rapid accumulation, facies stacking, and synsedimentary
from a relatively dense grid of seismic lines (Figure 18) deformation adjacent to basin-bounding faults. Tupe and
(YPFB proprietary files). These paleovalleys erode deeply Patquia–De la Cuesta stratigraphy gradually onlapped
into the Machareti section and have themselves been the interbasin highs as the earlier transtensional basins
eroded by pre-Cretaceous structural inversion. The were amalgamated in a single, broadly subsiding basin
Escarpment paleovalleys are typically up to 500 m deep (Figure 17).
and consist internally of stacked channel sandstones and This ancient system of crustal strike-slip faults
mudstones. There were at least five episodes of paleo- diverges northward where it combines with the east-
valley incision and aggradation suggesting intermittent west Boomerang Hills–Chiquitanas suture to form the
base level changes. deep Chaco basin of southern Bolivia (Figure 18). The
In the upper Mandiyuti, the San Telmo Formation Chaco salient of the Bolivian Andes reflects this thick
contains poorly bedded diamictites with striated clasts, sedimentary pile.
shallow cross-bedded fluvial sandstones, and varved In the Boomerang area, seismic data shows exten-
lacustrine shales with plant fossils. Pollen assemblages in sional reactivation of basement structures during the
the varved shales of the Cañada Honda well indicate a middle Paleozoic. This may coincide with Early
nonmarine aquatic environment of Stephanian age Carboniferous Tupambi deposition.
equivalent to the Tupe of the Paganzo basin (Azcuy, The Machareti and Mandiyuti groups of Bolivia and
1979; W. A. M. Jenkins, 1991, personal communication). northernmost Argentina are stratigraphically equivalent
We infer a periglacial outwash plain paleogeography. to the Guandacol and Tupe supersequences, respec-
tively. The paleogeographic setting of stacked braided
alluvial fans and lacustrine sediments in the Machareti
DISCUSSION Group, together with evidence of rapid subsidence,
suggest a tectonic setting similar to that of the Guandacol
Fieldwork and subsurface data provide convincing Paganzo basin. We believe that extensional or transten-
evidence that the Permian–Carboniferous Paganzo basin sional processes also affected the Machareti. However,
started its history as a suite of isolated but structurally the Tarija and Chaco depocenters were shallower but
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 299

more widespread, reflecting the northward-diverging REFERENCES CITED


shear system (Figures 18, 19). Paleocurrent analysis
shows that these depocenters were filled from the east Aceñolaza, F. G., and A. J. Toselli, 1976, Consideraciones
and west. estratigráficas y tectónicas sobre el Paleozoico inferior del
The system of paleovalleys mapped seismically in the noroeste Argentino: Segundo Congreso Latinoamericano
Chaco basin is believed to have drained into transten- de Geología, v. 2, p. 754–764.
Aceñolaza, F. G., and M. M. Vergel, 1987, Hallazgo del
sional basins. Its orientation matches the extensional
Pérmico superior fosilífero en el sistema del Famatina:
trend. This interpretation suggests that the repetitive Décimo Congreso Geológico Argentino, Actas 3,
base level adjustments reflected in paleovalley formation p. 125–129.
and aggradation may have a tectonic explanation. Alvarez, L. A., F. Fernandez-Seveso, M. A. Perez, and N. D.
Sediment was apparently derived from the east and Bollatti, 1990, Estratigrafía de la Cuenca Saliniana: Décimo
southeast along the Chaco Boreal high (Sempere, 1995). Primer Congreso Geológico Argentino, Actas 2, p. 145–148.
Like the Tupe, the Mandiyuti cover is regionally more Amos, A. J., 1972, Las cuencas carbónicas y pérmicas de la
widespread. Sempere (1995) notes that the Upper Argentina: Anales Academia Brasileira de Ciencias, v. 44,
Carboniferous–Permian stratigraphy onlaps older paleo- p. 21–36.
relief, much like the Tupe–Patquía of northwestern Amos, A. J.,and E. O. Rolleri, 1965, El Carbónico marino en el
Valle Calingasta-Uspallata (San Juan-Mendoza): Boletín
Argentina. A prominent transgression occurred in late
Informaciones Petroleras, v. 368, p. 50–71.
Kungurian–Kazanian time, during which the Vitiacua Andreis, R. R., and O. G. Arrondo, 1974, Acerca de la discor-
calcareous mudstones were deposited. The Vitiacua has dancia angular entre las Formaciones Guandacol y Tupe
high TOC values and is equivalent to the upper en la Sierra de Maz (Provincia de La Rioja): Ameghiniana,
Patquía–De la Cuesta and Irati (Brazil) formations. Revista Asociación Paleontológica Argentina, Buenos
Although oil-prone, the Vitiacua has so far attracted little Aires, v. 11, p. 373–378.
exploration interest. Archangelsky, S., 1986, El sistema carbonífero en la República
This tectonostratigraphic interpretation highlights Argentina (Síntesis): UNESCO Proyecto 211, 359 p.
several issues which have yet to be addressed: Azcuy, C. L., 1979, A review of the early Gondwana paly-
nology of Argentina and South America: IV International
Palynological Conference, Lucknow 2, p. 177–208.
• The Paganzo–Ischigualasto basin complex has
Azcuy, C. L., and J. R. Morelli, 1970, Geología de la comarca
evolved through several episodes of middle Paganzo-Amaná. El Grupo Paganzo. Formaciones que lo
Paleozoic strike-slip movement, Mesozoic componen y sus relaciones: Revista Asociación Geológica
extension and inversion, and late Tertiary contrac- Argentina v. 25-4, p. 405–429.
tion that have created the present tilt-block basin Baldis, B. A. J., M. Beresi, O. Bordonaro, and A. Vaca, 1982,
province (e.g., Fielding and Jordan, 1988). The Síntesis evolutiva de la Precordillera Argentina: Quinto
possibility of multiple source rock intervals Congreso Latinoamericano de Geología, Actas 4,
suggests the need to unravel this structural p. 399–495.
complexity as a guide to source rock distribution Baldis, B. A. J., and G. A. Chebli, 1969, Estructura profunda
and maturation. del área de la Precordillera Sanjuanina: Cuarta Jornada
Geológica Argentina, v. 1, p. 47–65.
• In Bolivia, a detailed biostratigraphic study of the
Bally, A. W., and S. Snelson, 1980, Realms of subsidence, in A.
Carboniferous section is needed. The dynamics of a D. Miall, ed., Facts and principles of world petroleum
transtensional basin system cast doubt on existing occurrence: Canadian Society of Petroleum Geologists
stratigraphic correlations. Memoir 6, p. 9–94.
• There is a striking resemblance of the Chaco salient Bond, G. C., and M. A. Kominz, 1984, Construction of tectonic
of the Bolivian Andes and the Mackenzie subsidence curves for the early Paleozoic miogeocline,
Mountains of northern Canada (Eisbacher, southern Canadian Rocky Mountains: implications for
1985).The Chaco and Mackenzie salients are both subsidence mechanism, age of break up, and crustal
associated with thick Proterozoic and Paleozoic thinning: GSA Bulletin, v. 95, p. 155–173.
depositories, and each has detached and trans- Caballé, M. F., 1990, Magmatismo permo-triásico al oeste de
Calingasta, Cordillera Frontal de San Juan, Argentina:
ported Carboniferous basins of probable pull-apart
Décimoprimer Congreso Geológico Argentino, Actas 1,
origin (Machareti and Antler) (H. R. Balkwill, 1994, p. 28–31.
personal communication). A comparison of these Caminos, R., 1985, El magmatismo neopaleozoico en la
two thrust belts would be fruitful. Argentina. Síntesis y principales problemas: UNESCO,
Proyecto 211, Encuentro anual del grupo Argentino de
trabajo, p. 1–15.
Acknowledgments We thank L. Alvarez, P. Aukes, C. Cingolani, C. A., R. Varela, E. Morel, O. Schauer, O. G.
Azcuy, I. E. Brisson, M. Cirbian, T. E. Jordan, O. Lopez- Arrondo, 1990, Aportes bioestratigráficos en el Devónico-
Gamundi, O. Lopez Paulsen, M. A. Perez, V. A. Ramos, K. Carbónico del sector septentrional de la Sierra de la
Raskin, and M. A. Uliana for helpful discussions. M. Arguijo, Punilla, provincia de La Rioja: Décimo Congreso
Geológico Argentino, Actas 2, p. 207–210.
O. Decastelli, and M. Distefano helped with the geochemistry.
Christie-Blick, N. and K. T. Biddle, 1985, Deformation and
Drafting was undertaken by D. Betty and S. Bond and typing basin formation along strike-slip faults, in K. T. Biddle and
by K. Bojarski. The paper was reviewed by C. J. Schmidt, H. J. N. Christie-Blick, eds., Strike-slip deformation, basin
Belotti, and H. J. Welsink. Finally, we thank the exploration formation, and sedimentation: SEPM Special Publication
management of YPF SA for support and permission to publish. 37, p. 1–34.
300 Fernandez-Seveso and Tankard

Dickins, J. M., 1993, Climate of the Late Devonian to Triassic: Mann, P., M. Hempton, D. Bradley, and K. Burke, 1983,
Palaeogeography, Palaeoclimatology, Palaeoecology, v. Development of pull-apart basins: Journal of Geology, v.
100, p. 89–84. 91, p. 529–554.
Eisbacher, G. H., 1985, Pericollisional strike-slip faults and Meyer L., 1991, Central Los Angeles basin, subsidence and
synorogenic basins, Canadian Cordillera, in K. T. Biddle thermal applications for tectonic evolution, in K. T. Biddle,
and N. Christie-Blick, eds., Strike-slip deformation, basin ed., Active margin basins: AAPG Memoir 52, p. 185–196.
formation, and sedimentation: SEPM Special Publication Milana, J. P., and F. Bercowski, 1990, Facies y geometría de
37, p. 265–282. depósitos glaciales en un paleovalle carbonífero de
Eyles, N., 1993, Earth’s glacial records and its tectonic setting: precordillera central, San Juan, Argentina: Tercera Reunión
Earth-Science Reviews, v. 35, p. 1–248. Argentina de Sedimentología, p. 199–204.
Eyles, N., G. Gonzalez-Bonorino, A. B. Franca, and C. H. Milana, J. P., F. Bercowski, and R. R. Lech, 1987, Análisis de la
Eyles, 1995, Hydrocarbon-bearing late Paleozoic glaciated secuencia marino-continental neopaleozoica, en la región
basins of southern South America, in A. J. Tankard, R. del río San Juan, precordillera central, Argentina: Décimo
Suarez, and H. J. Welsink, Petroleum basins of South Congreso Geológico Argentino, Actas 3, p. 113–116.
America: AAPG Memoir 62, this volume. Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979,
Fernandez-Seveso, F., L. A. Alvarez, M. A. Perez, and E. Sierras Subandinas, in J. C. M. Turner, ed., Geología
Rodriguez, 1988, Sistemas depositacionales de las sedi- regional Argentina: Segundo Simposio, Cordoba,
mentitas eopérmicas de Bajo de Veliz, Tasacuna y A° Academia Nacional de Ciencias, v. 1, p. 95–138.
Totoral: Segunda Reunión Argentina de Sedimentología, Mon, R., and J. A. Salfity, 1995, Tectonic evolution of the
p. 90–94. Andes of northern Argentina,in A. J. Tankard, R. Suarez,
Fernandez-Seveso, F., M. A. Perez, and L. A. Alvarez, 1990, and H. J. Welsink, Petroleum basins of South America:
Análisis estratigráfico del ámbito occidental de la cuenca AAPG Memoir 62, this volume.
de Paganzo, en el rango de grandes ciclos depositacionales: Mpodozis, C., and S. M. Kay, 1990, Provincias magmáticas
Decimoprimer Congreso Geológico Argentino, Acta 2, ácidas y evolución tectónica de Gondwana, Andes
p. 77–80. Chilenos (28–31° S): Revista Geológica de Chile, v. 17,
Fernandez-Seveso F., M. A. Perez, and I. E. Brisson, 1991, p. 152–180.
Estratigrafía y potencial generador de hidrocarburos en los Mpodozis, C., and V. A. Ramos, 1989, The Andes of Chile and
Boisones de Fiambalá y Chaschuil: YPF proprietary report, Argentina, in G. E Ericksen, M. T. Cañas Pinochet, and J. A.
182 p. Reinemud, eds, Geology of the Andes and its relations to
Fernandez-Seveso, F., M. A. Perez, I. E. Brisson, and L. A. hydrocarbon and mineral resources: Circum-Pacific
Alvarez, 1993, Sequence stratigraphy and tectonic analysis Council for Energy and Mineral Resources, Earth Science
of the Paganzo basin, western Argentina: XII International Series, v. 11, p. 59–90.
Congress on Carboniferous–Permian, v. 2, p. 223–260. Mutti, E., 1985, Turbidite systems and their relations to depo-
Fielding, E. J., and T. E. Jordan, 1988, Active deformation at sitional sequence, in G. Zuffa, ed., Provenance of arenites:
the boundary between the Precordillera and Sierras NATO-ASI Series, p. 65–93.
Pampeanas, Argentina, and comparison with ancient Nilsen, T. H., and R. J. McLaughlin, 1985, Comparision of
Rocky Mountain deformation, in C. J. Schmidt and W. J. tectonic framework and depositional patterns of the
Perry, Jr., eds., Interaction of Rocky Mountain foreland and Hornelen strike-slip basin of Norway and the Ridge and
Cordillera thrust belt: GSA Memoir 171, p. 143–163. Little Sulphur Creek strike-slip basins of California, in K. T.
França, A. B., E. J. Milani, R. L. Schneider, O. Lopez, M. Lopez, Biddle and N. Christie-Blick, eds., Strike-slip deformation,
R. Suarez, H. de Santa Ana, F. Wiens, O. Ferreiro, E. A. basin formation, and sedimentation: SEPM Special Publi-
Rossello et al., 1995, Phanerozoic correlation in southern cation 37, p. 79–103.
South America,in A. J. Tankard, R. Suarez, and H. J. Perez, M. A., F. Fernandez-Seveso, L. A. Alvarez, and I. E.
Welsink, Petroleum basins of South America: AAPG Brisson, 1993, Análisis ambiental y estratigráfico del Paleo-
Memoir 62, this volume. zoico Superior en el área del anticlinal de Huaco. San Juan,
Gonzalez, C. R., and G. E. Bossi, 1986, Los depósitos Argentina: XII International Congress on
carbónicos al Oeste de Jagüel, La Rioja: Cuarto Congreso Carboniferous–Permian, v. 2, p. 297–318.
Argentino Paleontología y Bioestratigrafía, Acta 1, Polansky, J., 1970, Carbónico y Pérmico de la Argentina:
p. 231–236. Buenos Aires, Eudeba, 216 p.
Gordillo, C. E., and A. N. Lencinas, 1979, Sierras Pampeanas Ramos, E. D., and V. A. Ramos, 1979, Los ciclos magmáticos
de Córdoba y San Luis. Geología Regional Argentina: de la República Argentina: Séptimo Congreso Geológico
Academia Nacional de Ciencias de Córdoba, v. 1, Argentino, v. 1, p. 771–786.
p. 577–650. Ramos, V. A., 1988a, Late Proterozoic–early Paleozoic of
Limarino, C. O., and L. A. Spalletti, 1986, Eolian Permian South America: a collisional history: Episodes, v. 11,
deposits in west and northwest Argentina: Sedimentary p. 168–174.
Geology, v. 49, p. 109–127. Ramos, V. A., 1988b, The tectonics of central Andes; 30°–33° S
Lopez-Gamundi, O., L. A. Alvarez, R. R. Andreis, I. Espejo, F. latitude, in S. Clark and D. Burchfiel, eds., Processes in
Fernandez–Seveso, D. A. Kokogián, L. Legarreta, C. O. continental lithospheric deformation: GSA Special Paper
Limarino, and H. L. Sessarego, 1989, Cuencas intermon- 218, p. 31–54.
tanas neopaleozoicas: Décimo Congreso Geológico Ramos, V. A., T. E. Jordan, R. W. Allmendinger, C. Mpodozis,
Argentino, Simposio de Cuencas Sedimentarias, S. M. Kay, J. M. Cortes, and M. A. Palma, 1986, Paleozoic
p. 123–167. terranes of the central Argentine–Chilean Andes:
Llambias, E. J., and A. M. Sato, 1990, El batolito de Colangüil Tectonics, v. 5, p. 855–880.
(29–31° S), Cordillera Frontal de Argentina, estructura y Rapalini, A. E., 1989, Estudio paleomagnético del volcanismo
marco tectónico: Revista Geológica de Chile, v. 17, permo-triásico de la región andina de la República
p. 89–108. Argentina. Consecuencias tectónicas y geodinámicas:
Tectonics and Stratigraphy, Late Paleozoic Paganzo Basin of Western Argentina 301

Ph.D. dissertation, University of Buenos Aires, Argentina, Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M.
278 p. Turic, A. B. Franca, E. J. Milani, B. B. de Brito Neves, N.
Reyes, F. C., 1972, On the Carboniferous and Permian of Eyles, H. de Santa Ana, et al., 1995, Structural and tectonic
Bolivia and northwestern Argentina: Anales da Academia controls of basin evolution in southwestern Gondwana
Brasileira de Ciencias, v. 44, p. 261–277. during the Phanerozoic, in A. J. Tankard, R. Suarez, and H.
Rivano, S., and P. Sepúlveda, 1985, Las calizas de la J. Welsink, Petroleum basins of South America: AAPG
Formación Huentelauquén: depósitos de aguas templadas Memoir 62, this volume.
a frías en el carbonífero superior-pérmico inferior: Revista Trujillo Ikeda, H., 1989, Nuevo hallazgo de fósiles de la
Geológica de Chile, no. 25-26, p. 29–38. formación Taiguati en la Serranía Caipipendi, Santa Cruz,
Rodriguez, E., and F. Fernandez-Seveso, 1988, Origen y Bolivia: Revista Técnica de Yacimientos Petroliferos
estructura de las ritmitas del lago eopermico Bajo de Veliz, Fiscales Bolivianos, v. 10, p. 7–11.
San Luis: Segunda Reunión de Sedimentología, p. 227–231. Vásquez, J., R. Gorroño, and E. Ivorra, 1981, El Paleozoico
Russo, A., S. Archangelsky, R. R. Andreis, and A. J. Cuerda, Superior de las provincias de San Juan y La Rioja: Revista
1986, Cuenca Chacoparanaense, in S. Archangelsky, ed., El de la Asociación Geológica Argentina, v. 36, p. 89–98.
sistema Carbonífero en la Argentina: UNESCO, Proyecto Villar, H. J., and O. R. Lopez-Gamundi, 1993, Carbones y
211, p. 183–198. pelitas carbonosas del carbonífero de la cuenca de
Ryan, H. F., and P. J. Coleman, 1992, Composite Paganzo: contexto litofacial y potencial generador de
transform–convergent plate boundaries: description and hidrocarburos: Décimosegundo Congreso Geológico
discussion: Marine and Petroleum Geology, v. 9, p. 89–97. Argentino y Segundo Congreso de Exploración de Hidro-
Salfity, J. A., and S. A. Gorustovich, 1983, Paleogeografía de la carburos, v. 1, p. 375–381.
cuenca del Grupo Paganzo: Revista de la Asociación Visser, J. N. J., 1991, Geography and climatology of the Late
Geológica Argentina, v. 37, p. 437–453. Carboniferous to Jurassic Karoo basin in southwestern
Salfity, J. A., R. Omarini, B. Balbis, and W. J. Gutiérrez, 1975, Gondwana: Annals of the South African Museum, v. 99,
Consideraciones sobre la evolución geológica del Precám- p. 415–431.
brico y Paleozoico del norte Argentino: Segunda Congreso Williams, K. E., 1995, Tectonic subsidence analysis and a
Iberoamericano de Geológica Económica, Buenos Aires, Paleozoic paleogeography of Gondwana, in A. J. Tankard,
Actas 4, p. 341–361. R. Suarez, and H. J. Welsink, Petroleum basins of South
Scalabrini-Ortiz, J., 1973, El Carbónico de la Precordillera America: AAPG Memoir 62, this volume.
Argentina al norte del río Jachal: Quinto Congreso Willner, A. P., U. S. Lottner, and H. Miller, 1987, Early
Geológico Argentino, Actas 3, p. 387–401. Paleozoic structural development in the NW Argentine
Scholz, C. A., B. R. Rosendahl, and D. L. Scott, 1990, Develop- basement of the Andes and its implication for geodynamic
ment of coarse-grained facies in lacustrine rift basins: reconstructions, in G. D. McKenzie, ed., Gondwana six:
examples from East Africa: Geology, v. 18, p. 140–144. structure, tectonics, and geophysics: American Geophys-
Scotese, C. R., and S. F. Barrett, 1990, Gondwana’s movement ical Union, Geophysical Monograph 40, p. 229–239.
over the South Pole during the Palaeozoic: evidence from Zalan, P. V., S. Wolff, J. C. J. Conceiçao, M. A. M. Astolfi, I. S.
lithological indicators of climate, in W. S. McKerrow and Vieira, V. T. Appi, and O. A. Zanotto, 1987, Tectônica e
C. R. Scotese, eds., Palaeozoic palaeogeography and sedimentacao da Bacia do Paraná: III Simposium Sul-
biogeography: Geological Society of London Memoir 12, Brasileiro de Geologia, v. 1, p. 441–447.
p. 75–85.
Sempere, T., 1995, Phanerozoic evolution of Bolivia and
adjacent regions, in A. J. Tankard, R. Suarez, and H. J.
Welsink, Petroleum basins of South America: AAPG
Memoir 62, this volume. Authors’ Mailing Addresses
Sempere, T., E. Aguilera, J. Doubinger, P. Janvier, J. Lobo, J. Fernando Fernandez–Seveso
Oller, and S. Wenz, 1992, La Formation de Vitiacua YPF SA
(Permien moyen à supérieur–Trias? inférieur, Bolivie du Av. Roque Saenz Peña 777
Sud): stratigraphie, palynologie et paléontologie: Neues
Jahrbuch für Geologie und Paläontologie, Abhandlungen,
1364 Buenos Aires
v. 185, p. 239–253. Argentina
Smith, A. G., A. Hurley, and J. C. Briden, 1981, Phanerozoic
paleocontinental world maps: Cambridge, New Jersey, Anthony J. Tankard
Cambridge University Press, 102 p. Tankard Enterprises
Starck, D., E. Gallardo, and A. E. Schultz, 1992, La cuenca de P.O. Box 81002
Tarija: estratigrafia de la porción Argentina: Boletin de Calgary, Alberta T2J 7C9
Informaciones Petroleras, Tercera Epoca, no. 30, p. 2–14. Canada
Mesozoic Rifts

QUEBRADA DE HUMAHUACA, Jujuy province, Argentina.


Orange-colored rift deposits of the Upper Cretaceous Salta Group are
unconformable on purple Cambrian strata and are overthrust by green
Ordovician rocks.

Edgar Ortiz, 1994, watercolor, 30 × 23 cm


Structural Inversion of a Cretaceous Rift Basin,
Southern Altiplano, Bolivia

H. J. Welsink E. Martinez
Perez Companc O. Aranibar
Neuquén, Argentina J. Jarandilla
Yacimientos Petrolíferos Fiscales Bolivianos
Santa Cruz de la Sierra, Bolivia

Abstract

T he southern Altiplano rift basin forms part of the Cretaceous rift system that extends from Peru to north-
western Argentina. Seismic and potential field data suggest that basin formation was controlled by a
preexisting structural grain consisting of northwest-southeast and northeast-southwest tectonic lineaments
within a Precambrian–Paleozoic basement. Variable extension was achieved through transfer fault zones that
coincide with northwest-southeast trending lineaments. Rift subsidence can be divided into an early rift (Berri-
asian–Cenomanian) and a late rift episode (Turonian–Campanian) with a total accumulation of up to 1000 m of
sandstones and shales. A period of tectonic quiescence followed that resulted in a regional sag basin character-
ized by thin-bedded, calcareous lacustrine deposits. This sequence includes the oil-prone shales and reservoir
carbonates of the El Molino Formation, which has been correlated with the hydrocarbon-producing Yacoraite
Formation in northwestern Argentina.
At the beginning of the Tertiary, rift-related subsidence had ceased and was gradually replaced by subsi-
dence resulting from Andean compression. An Eocene unconformity marks the subtle change to this new
episode of basin formation. This was followed by a major Oligocene unconformity that characterizes the onset
of increased subsidence rates related to the emplacement of large thrust sheets of the Eastern Cordillera. More
than 4 km of synorogenic sediments accumulated in the adjacent Altiplano foredeep. Inversion of the Creta-
ceous rift structures took place during this compressional phase. In the southern Altiplano, this inversion
resulted in hydrocarbon traps similar to those in the rift basins of northwestern Argentina.

Resumen

L a cuenca de rift del Altiplano del sur es parte del sistema extensional cretácico andino, extendiéndose
desde Perú hasta el noroeste argentino. Datos sísmicos, gravimétricos y magnéticos sugieren que la
evolución de la cuenca fue controlada por un marco estructural preexistente que consiste en lineamientos
tectónicos de rumbo noroeste-sureste y noreste-suroeste dentro de un basamento precámbrico-paleozoico. La
extensión variable fue conseguida por zonas de fallas de transferencia que coinciden con lineamientos de
rumbo noroeste-sureste. La subsidencia del rift es divida en un episodio de rift temprano y tardío, con una
acumulación total de hasta 1000 m de areniscas y pelitas. Continuó un período de quietud tectónica que resultó
en una cuenca de subsidencia regional caracterizada por depósitos lacustres calcáreos en capas delgadas. Estos
sedimentos incluyen las pelitas generadoras de petróleo y calizas reservorio de la Formación El Molino, la cual
fue correlacionada con la Formación Yacoraite, productiva de hidrocarburos en el noroeste argentino.
En el inicio del Terciario, la subsidencia del rift había concluido siendo reemplazada gradualmente por
subsidencia resultante de la compresión andina. Una discordancia eocena marca el cambio sutil a este nuevo
episodio de formación de cuenca, seguido por una discordancia oligocena que caracteriza el incremento en las
velocidades de subsidencia relacionado al emplazamiento de láminas de corrimiento en la Cordillera Oriental.
Mas de 4 km de sedimentos sinorogénicos fueron acumulados en la adyacente antefosa del Altiplano. La
inversión de las estructuras cretácicas de rift ocurrió durante esta fase compresiva. En el Altiplano del sur esta
inversión resultó en trampas de hidrocarburos similares a las documentadas en las cuencas de rift del noroeste
argentino.

Welsink, H. J., E. Martinez,. O. Aranibar, and J. Jarandilla, 1995, Structural inversion of a 305
Cretaceous rift basin, southern Altiplano, Bolivia, in A. J. Tankard, R. Suárez S., and H.
J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 305–324.
306 Welsink et al.

INTRODUCTION
During the Cretaceous, several basins developed
along the western margin of South America, now a
major hydrocarbon-producing province (Macellari,
1988). Bolivia forms an exception, as production pri-
marily comes from Paleozoic reservoirs. As a conse-
quence, the southern Altiplano is an area considered to
be potentially prospective but yet without commercial
hydrocarbon production. The presence of a gas seep in
Corregidores, an oil seep in Colchani, and impregnations
of solid petroleum in Casira (Figure 1) have allowed us
to define possible Cretaceous and Devonian source rocks
(Edman et al., 1989).
The present study was undertaken to reevaluate the
hydrocarbon potential of the Vilque area in the southern
Altiplano (Figure 1) and to document inversion of rift-
related structures along the northern flank of the San
Pablo high that until recently had not been considered
part of the Andean rift basin (Figure 2). Comparison of
these structures with those of the oil-producing rift
basins of northwestern Argentina creates opportunities
favorable for exploration in the Altiplano.

REGIONAL GEOLOGIC FRAMEWORK


Within the Andean orogen, the Altiplano-Puna
plateau forms a geologic province that dominates the
central Andes of southern Peru, Bolivia, northern Chile,
and northern Argentina (Figure 1). Located between the
Eastern and Western Cordillera, it covers about 140,000
km2 from southwestern Peru to northwestern Argentina.
In Bolivia, the Altiplano covers about 100,000 km2, its
surface being a plateau, 3800–4000 m above sea level,
that is characterized by Quaternary fill of fluvial,
fluvioglacial, and lacustrine sediments. In the southern
Altiplano, the salt flats (Salar) of Uyuni and Coipasa
comprise evaporitic lithium-bearing sediments that are
of economic interest. Highlands within the Altiplano
consist of sedimentary rocks ranging in age from
Precambrian to Cenozoic. Cenozoic volcanism is ex-
pressed as intrusives, lava flows, and extensive ignim-
britic flows, especially in the central Altiplano.
From a structural viewpoint, the Altiplano forms a
contractional tectonic province that was deformed and Figure 1—Location of Altiplano with respect to the Andean
uplifted in the Tertiary as a result of the Andean front of Eastern Cordillera and the volcanic belt of Western
orogeny. The mechanisms of crustal thickening and Cordillera. Major tectonic lineaments divide the Altiplano
uplift remain poorly understood and are based on into northern, central, and southern segments. The study
models of magmatic underplating (Thorpe et al., 1981) area (dotted) within the southern segment is confined by
or, more recently, crustal shortening (Allmendinger et al., the Uyuni-Kenyani fault zone to the west, the Andean
1983; Isacks, 1988; Sheffels, 1988). Interpretations of the thrust front to the east, and the San Pablo high to the
Cretaceous succession reveal a preexisting extensional south. Alota and Pululus are outcropping transpressional
tectonic province that extended from Peru to north- structures composed of Ordovician rocks.
western Argentina, what is now known as the Creta-
ceous Andean basin (Moreno, 1970; Reyes, 1972; Reyes porated in thrust sheets that terminate at the southern rift
and Salfity, 1973; Cherroni, 1977) (Figure 2). border (Figure 2). Away from this front, only incipient
Beginning in the late Oligocene, the Andean orogeny inversion of rift-related normal faults took place
overprinted the extensional structures as they became (Bianucci et al., 1982).
involved to varying degrees in the Andean deformation. Major tectonic elements that characterize the various
The rift basins east of the Andean thrust front are incor- structural domains of the Altiplano have been identified
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 307

Figure 2—Northwestern Argen-


tinian rift system and its
extension into neighboring
countries. Major tectonic linea-
ments localize extension and
control the distribution of Early
Cretaceous volcanism.
(Modified after Cherroni, 1977;
Salfity,1985; Marquillas and
Salfity, 1988.)

(Aranibar and Martinez, 1990; Cady and Wise, 1992). and during this period acquired much of the geologic
Figure 1 shows a simplified division into three segments and geophysical data. Surface mapping covered 40,000
separated by major northwest-southeast tectonic linea- km2, and 8900 km of seismic reflection data were shot.
ments. The southern segment includes the southern YPFB also collected gravity data and was responsible for
Altiplano basin, which is constrained by three major the registration of an aeromagnetic survey in 1969.
tectonic elements. To the west and north, the Uyuni- Results of these studies are discussed in several propri-
Kenyani fault zone forms a linear suite (or belt) of trans- etary YPFB reports that were used in this project.
pressional structures that contain Ordovician, Silurian, Only five wells have been drilled (Figure 1), all of
Devonian, Cretaceous, and Tertiary sediments. To the which were dry. However, the Vilque well in the
east, westward-verging thrust sheets expose a similar southern Altiplano had dry gas shows in the Neogene–
stratigraphy. This fold and thrust belt is a westward- Oligocene San Vicente Formation. Based on these disap-
verging backthrust belt of the main Andean thrust pointing well results, this first phase of exploration was
system (Roeder, 1988). The San Pablo high and the terminated. Interest in the area was renewed in 1986 with
volcanic belt of the Western Cordillera form the southern reinterpretation of seismic data of the southern Altiplano
margin of the study area. and digital conversion of the gravity data. Most recently,
an extensive aeromagnetic survey was acquired in 1989.
Exploration History Preliminary results of these recent studies are given by
Aranibar and Martinez (1990) and Yalamanchili et al.
Yacimientos Petrolíferos Fiscales Bolivianos (YPFB), (1991). The present paper presents an interpretation
the Bolivian national oil company, executed the first resulting from the compilation and integration of all
phase of Altiplano exploration between 1960 and 1976, available data.
308 Welsink et al.

At present, the Altiplano basins do not produce any


commercial hydrocarbons. However, three factors
suggest that these basins are prospective: (1) about 45,000
tons of oil have been produced in Pirín at the southern
end of Lake Titicaca in Peru, (2) nearby producing fields
in northwestern Argentina source the oil from the Upper
Cretaceous Yacoraite Formation, and (3) there are oil and
gas seeps in the Altiplano. We believe that the Altiplano
remains a potentially attractive hydrocarbon play area.
Recently, the Upper Cretaceous El Molino Formation has
been identified as a potential source rock in the Altiplano
basins, while hydrocarbons from seeps have been corre-
lated with Devonian and Cretaceous source intervals
(Edman et al., 1989; M. Cirbián, 1992, personal communi-
cation).

METHODOLOGY AND DATABASE


The present study addresses available geologic and
geophysical data, including geologic surface maps,
gravity, aeromagnetics, reflection seismic, and well infor-
mation. These have been integrated and interpreted. Our
approach has involved the evaluation and compilation of
available information, followed by preliminary seismic
interpretations and integration with the gravity,
magnetic, and field data. Preliminary interpretations
include the recognition of major faults, seismic strati-
graphic sequences, and their geometries. The resulting
tectonic elements map was subsequently used as a Figure 3—Main features of the southern Altiplano and
framework for more detailed mapping of seismic reflec- summary of available information used in the interpretation
tions and the construction of isopach maps to establish (Tupiza-Casira area excluded). Dotted area shows
the extensional architecture of the southern Altiplano coverage of geologic maps.

Table 1—Stratigraphy of the Southern Altiplano

Formation Thickness Description


Tertiary
Chocaya 52–950+ m Continental coarse-grained clastics and volcanic clastics, rapid facies changes; shales
and light gray tuffs
San Vicente 100–3600 m Alluvial system of sandstones and conglomerates grading into shales and tuffs in the
basin center. Interbedded salt
Potoco 35–2798 m Red fluviolacustrine sandstones, conglomeratic sandstones, and shales
Cayara 7–60 m Reworked fluviolacustrine sandstones
Sta. Lucia 30–90 m Lacustrine marls and mudstones with oxidation zones and paleosoils indicating exposure

Cretaceous
El Molino 38–350 m Basal conglomerates and sandstones followed by stacked shallowing-upward sequences
comprising thin-bedded calcareous lacustrine deposits; black marls
Chaunaca 64–400 m Lacustrine siltstones and shales, thin-bedded basal limestones, fossiliferous
Aroifilla 15–560 m Fluviolacustrine sandstones and shaly siltstones, evaporitic levels

Devonian
Vila-Vila 920 m Shallow marine arkosic sandstones

Silurian
Catavi 1600–3530 m Shallow marine shelf sandstones
Uncía 100–195 m Dark gray mudstones, sandstones interbedded with graptolite-bearing layers
Llallagua 60–350 m Packages of quartzitic sandstones and rhythmic dark gray mudstones; turbidites and
submarine fans
Cancañiri 100–540 m Diamictites, debris flows and slope deposits; heterogeneous glaciomarine deposits
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 309

basin. Finally, the hydrocarbon prospectivity was


evaluated by comparing the tectonostratigraphic
framework and inversion with those of the rift basins of
northwestern Argentina.
The geologic data used in this study were obtained
between 1960 and 1976, including geologic maps,
measured stratigraphic sections, and well information
(Figure 3). The field data of the Casira and Tupiza areas
are not included in the present study. The geologic maps
were generated by various YPFB field parties at original
1:20,000 and 1:50,000 scales; we have compiled these into
a map at a scale of 1:100,000 to obtain compatible inter-
pretations. This scale allows the identification of the
principal features without loss of detail and also corre-
sponds to a level of investigation that matches the
present seismic grid.
The measured sections were reinterpreted in terms of
major unconformity-bounded sequences and associated
facies changes before being integrated into the regional
seismic stratigraphic framework. The present study
includes 24 measured sections in the southern Altiplano
(Figure 3), representing a Paleozoic–Tertiary strati-
graphic column (Table 1, Figure 4).
Only one well is present in the southern Altiplano
(Figures 1, 3, 4). The Vilque-x1 was drilled in 1972–1973
to a total depth of 3559 m and was abandoned dry. The
well was drilled on the basis of structural relationships
expressed at the surface because no seismic data were
available at that time. Units encountered include the
Tertiary San Vicente and Potoco formations, the Creta-
ceous El Molino Formation, and undifferentiated Silurian
rocks at drilling depth (Figure 4). A thrust fault with 380
m of throw was encountered at 1835 m and repeated the
Potoco Formation. The well tested gas (80% nitrogen)
that flowed 1.4–1.9 mmcf per day from a zone below the
salt in the San Vicente Formation. Other tests recovered
water.
The geophysical database consists of gravity, aero-
magnetic, and reflection seismic data. The location of
these surveys is shown in Figure 3. The extensive gravity
database was acquired by YPFB between 1965 and 1974
and consists of 57,530 gravity stations covering approxi-
mately 80,000 km2 of the northern, central, and southern
Altiplano (see Cady and Wise, 1992, for an overview). Figure 4—Lithologic column of the Vilque well. (See
Figures 1 and 3 for location.)
The 0.5-km by 5-km grid in the southern Altiplano
contains about 15,000 stations (Figure 3). In 1987, the data
were reprocessed after digital conversion. The new Nearly 15,000 km of flight lines cover the southern
gravity maps form the basis of recent structural interpre- Altiplano. Yalamanchili et al. (1991) discuss acquisition
tations of the Altiplano (Aranibar and Martinez, 1990; and processing of these data.
Yalamanchili et al., 1991; Cady and Wise, 1992) and were In 1970, YPFB started the recording of 767 km of reflec-
used for this study after integration with other geophys- tion seismic lines in the southern Altiplano. These are
ical and geologic data sets. analog data and of low fold, and their quality is conse-
The first geophysical reconnaissance in the study area quently poor. In 1974–1975, YPFB acquired 1200 km of 48-
was acquired in 1967 when 17,000 km of flight lines of channel vibroseis data in the Salar de Uyuni area north of
aeromagnetic data were registered. Interpretation of the Uyuni-Kenyani fault zone. South of the Uyuni-
these data defined the major magnetic lineaments that Kenyani fault zone, the Vilque area was covered by 17
were attributed to structural boundaries between seismic lines totaling 705 km in a grid varying between 10
different geologic provinces. In 1989, a new high-resolu- km by 10 km and 5 km by 5 km (Figure 3). The data were
tion aeromagnetic survey was flown in the southern and processed to a final stack of 24-fold and reprocessed for
central Altiplano, consisting of 32,841 km of flight lines in this study. The quality of the lines is generally good
a 2-km (north-south) by 6-km (N 120˚ W long) grid. except for areas where subsurface salt is present.
310 Welsink et al.

Figure 5—Bouguer gravity


data with principal outcrop
trends mapped by YPFB. Solid
areas represent outcrops of
Cretaceous sediments of the
rift and sag subsidence
phases. The Bouguer values
were calculated using a
density of 2. 67 g/cm3. (YPFB
database was processed in
cooperation with Freie Univer-
sität Berlin, Germany.)

INTERPRETATION compatible with the seismic data that we used to


establish and correlate the seismic stratigraphic
Compilation of Geologic Maps sequences. We propose a tectonostratigraphic column
comprising five megasequences (Figures 6, 10). The most
The major structural trends and Cretaceous outcrop prominent unconformities are those at the base of the
locations are shown in Figure 5. The rocks that outcrop Cretaceous and at the base of the Tertiary San Vicente
are Ordovician, Silurian, Devonian, Cretaceous, and Formation. Between them are the rift and sag mega-
Tertiary in age. The Cretaceous outcrops (Figure 5) are sequence (4A and 4B), a transitional zone (5A), and the
no older than the Aroifilla Formation (Figure 6). Struc- foredeep stage (5B). This transition belongs to mega-
tural trends reflect anticlinal axes and thrust faults sequence 5 because it represents the initial phase of fine-
oriented in northeast-southwest and north-south direc- grained foredeep deposition and subdued topography
tions. The map shows the Uyuni-Kenyani fault zone as a (Potoco Formation), followed by coarse-grained molasse-
northeast-southwest trending linear suite of thrust faults type sediments (San Vicente Formation) (see Allen et al.,
attributed to a strike-slip fault zone. To the east, the San 1986).
Vicente thrust fault is conspicuous. This fault marks the This tectonostratigraphic column is an ideal represen-
deformation front of the westward-verging Altiplano tation of the subsurface based on available field data and
fold and thrust belt. seismic stratigraphic interpretations. The Lower Creta-
ceous succession (pre–Aroifilla Formation) does not
Stratigraphic Sequences outcrop in the immediate vicinity of the Vilque area, and
its presence is inferred from seismic stratigraphy. The
The compilation of measured field sections establishes stratigraphy reflects facies changes related to the position
and correlates unconformity-bounded sequences. As of the basin margin and shows the potential reservoir
major unconformities reflect the basic changes in basin and source rock intervals. Correlation with tectonic
tectonics, the intervening megasequences characterize events was achieved by structural–stratigraphic analysis
the evolution of the basin. This scale of analysis is more of the field data and seismic lines (Table 2).
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 311

Figure 6—Generalized
tectonostratigraphic chart of
the southern Altiplano
showing major unconformity-
bounded sequences derived
from seismic data and
regionally integrated with
measured field sections.
Arrows in last column
indicate fault activity: down,
normal; up, reverse
(inversion or thrusting). S,
source rocks; R, reservoir
rocks; white areas, shale;
solid triangles, diamictite.
Lithologic symbols same as
in Figure 4.

Gravity and Magnetic Trends mark the Uyuni-Kenyani fault zone. Internally, these
trends are dissected by northwest-southeast transverse
The gravity values of the southern Altiplano basin lineaments. A major change to north-south and
decrease progressively to the east (Figure 7); a –460 mgal northwest-southeast trends occurs north of Uyuni and is
minimum coincides with the 6000-m high mountains, characterized by a first-order transverse lineament that
suggesting that gravity values decrease with increasing marks the northern limit of the Uyuni-Kenyani fault
elevation (Cady and Wise, 1992). Depth to magnetic zone (Figures 1, 7).
basement exceeds 3000 m below sea level and increases Secondary ENE to WSW magnetic trends (Yalaman-
toward the central Altiplano (Yalamanchili et al., 1991). chili et al., 1991) are not resolved on the Bouguer gravity
Major lineaments were identified on the gravity and map. This is because gravity data only record Paleozoic
aeromagnetic data. Figures 7 and 8 show these linea- trends, while magnetic data also show structural trends
ments superimposed on the Bouguer gravity map of the in magnetic (Precambrian) basement. Similar discrepan-
southern Altiplano. In this area, the orientations of the cies are also present on the rifted margin of eastern
anomalies are mainly southwest-northeast and clearly Canada (Welsink et al., 1989b).
312 Welsink et al.

Table 2—Seismic Stratigraphic Sequences of the Southern Altiplano

Sequence Geometry Description


a
5B
Foredeep Westward thinning wedge Continental foredeep fill. Westward decrease in grain size and thickness
San Vicente suggest link with emplacement of thrust sheets in the Eastern
Cordillera. Seismically low to high amplitude and continuity and zones
without coherent reflections (salt); downlap indicates progradation
from the east.
5A
Transition Planar to slightly wedge Thermal sag to foredeep transition. Predominantly shale, but eastward
Potoco shaped increase of sandstone and conglomerate suggests an eastern source.
Cayara Moderate to high seismic amplitude. Good continuity decreases
southward. Onlaps Paleozoic highs.
4B
Sag Saucer shaped Regional sag reflecting maximum expansion of the Cretaceous basin
Sta. Lucía that onlaps and oversteps rift borders. Seismically poorly resolved
El Molino due to slight thickness, moderate amplitude and continuity.

4A
Late rift Wedge shaped Half-graben fill. Early rift succession reflects the decaying rift intensity
Chaunaca that culminates with deposition of Miraflores limestones. Continuing
Aroifilla expansion of the rift and late stage transition to thermal sag.
Early rift Seismically moderate to high amplitude and moderate continuity, with
Miraflores to Macha onlap onto Paleozoic basement.

3 Carboniferous is absent in the southern Altiplano.

1 and 2 Prerift basement, low to moderate amplitude without coherent


reflections.
aRepresents internal San Vicente reflection that forms top of a series of strong reflections (Figure 15).

Seismic Sequences spond to Paleozoic highs, gravity lows to Cretaceous


depocenters. As a result, the integration of the Bouguer
Preliminary interpretation of the seismic data is based gravity and seismic data has facilitated structural
on sequence geometries and the characteristics of reflec- mapping (Figure 14).
tion terminations. The sequence boundaries identified in Figures 5, 7, 8, and 9 show correlation between the
the Vilque well were tied to the major unconformities individual trends. Based on the seismic and field data,
recognized on the seismic data. Seismic line 2587-28 (see we interpret some lineaments on the gravity and
Figure 10) shows an example of the interpreted seismic magnetic maps as structural trends (anticlines) or discon-
sequences. Table 2 summarizes the characteristics of each tinuities (faults). The interpretation shown in Figure 14
megasequence. includes fabrics derived from the geologic map and the
gravity and aeromagnetic data (Figures 5, 7, 8).
Structural Styles
The structural elements define two main structural
styles (Figure 9). First, extensional structures are charac- BASIN ARCHITECTURE
terized by listric and planar normal faults of opposite Extensional basins generally consist of asymmetric
polarities (Figures 10, 11, 12, 13). Second, contractional half-grabens that are linked by either transfer faults
structures show reverse faults and their associated forced (Bally, 1982; Gibbs, 1984; Tankard and Welsink, 1987,
folds (Figures 12, 13). There appear to be no low-angle 1989) or complex accommodation zones (Scott and
thrust faults, although a shallow thrust within the Potoco Rosendahl, 1989). These transverse structures form an
Formation has been encountered in the Vilque well. integral part of the extensional system and accommodate
contrasting polarities and differential extension by
Discussion oblique-slip motion. The transverse offsets are mainly
parallel to the extension direction and often coincide
The structural interpretation of the seismic data is with lineaments (Welsink et al., 1989a).
shown in Figure 9, where it is superimposed on the Our integration of gravity and magnetic trends,
Bouguer gravity data. The structural elements on the surface geologic trends, and trends based on seismic data
seismic data match the gravity data: gravity highs corre- shows a typical rift basin geometry that is divided into
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 313

Figure 7—Bouguer gravity data


and major trends. Note change
in gravity trend from northeast
to northwest north of Uyuni.

three distinct areas (Figure 14). The transverse gravity recognize. Examples from the southern Altiplano,
and magnetic lineaments correspond to the boundaries however, show a progressive suite of structures that
of these structural segments. formed in response to low- and medium-grade inversion
North of the Vilque well, the basin has an unfaulted (Figures 10, 11, 12, 13, 15).
western margin and deepens toward the east. The central The sequence of structural inversion is well illustrated
segment is the deepest part of the basin; it is bounded to in the area between lines 2587-28 and 2585-28. Along line
the east by a Paleozoic high (Murusa) that is clearly 2587-28 (A–A', Figures 9, 10), there is no evidence of
defined by seismic, gravity, and field data. To the west, inversion and the seismic reflectors show only normal
listric basin-bounding faults have opposite polarities. The fault separations. Southward, line 25106-28 (B–B', Figures
southern segment is characterized by shallow depocen- 9, 11), shows the first indications of inversion at the rift
ters along the northern flank of the San Pablo high and borders. Subtle anticlines formed by forced folding are
by shallow intrusives interpreted from the aeromagnetic associated with inversion of the rift-bounding faults.
data (Yalamanchili et al., 1991). Line 2586-28 (C–C', Figures 9, 12) records a funda-
mental change in basin geometry across a transverse
lineament and also illustrates the development of a major
INVERSION pop-up structure bounded by reverse faults. The eastern
reverse fault is the most important. The extensional faults
Structural inversion involves the reactivation of an clearly cause normal stratigraphic separations, as well as
extensional fault system during contractional deforma- reverse separations after passing through the null point
tion. The process affects individual faults as well as entire (see Williams et al., 1989). This structure is shown on line
basins (Williams et al., 1989). This implies that an 2585-28 (D–D', Figures 9, 13) where, in what was then the
inverted structure requires a preceding extensional center of the rift, the western reverse fault formed the
phase. In highly deformed areas where destruction of the main fault. Seismic lines farther south do not show any
extensional geometries is severe, inversion is difficult to evidence of inversion and only a few rift structures.
314 Welsink et al.

Figure 8—Bouguer gravity data


with magnetic trends and
anomalies superimposed. The
higher amplitude magnetic
anomalies are associated with
shallow intrusives outside the
volcanic belt (Yalamanchili et
al., 1991). (Magnetic data
originate from aeromagnetic
survey acquired in 1989.)

Figures 14 and 15 show that the contractional struc- and gas seeps in the Altiplano mark the presence of a
tures are contained within the zone of rifting, thus mature Cretaceous source rock and an additional source
strongly supporting their correlation. The presence of rock that is believed to be of Paleozoic age (M. Cirbián,
Cretaceous rift deposits in the main thrust sheets of 1992, personal communication). This additional source
the Altiplano (e.g., Uyuni-Kenyani, Andamarca, and rock increases the prospectivity of the Altiplano
Sevaruyo) (Figure 1) indicates that they resulted from compared to northwestern Argentina, where Paleozoic
inversion of extensional faults within the Cretaceous rocks are progressively eroded from north to south.
Andean rift basin (Figure 2). In the southern Altiplano, the seismic data show
evidence of substantial amounts of normal faulting that
resulted in large fault traps and depocenters that were
HYDROCARBON PERSPECTIVES favorable to generation of hydrocarbons. Traps include
tilted normal fault blocks that were formed during Creta-
The development of extensional basins in the ceous extension and contractional structures due to
Altiplano and in northwestern Argentina was initiated at inversion or newly formed thrusts during the late
the beginning of the Cretaceous. Early and late rift stages Oligocene and later. We believe that inversion has locally
were succeeded by a Late Cretaceous regional sag basin destroyed the Cretaceous traps and caused remigration
characterized by thin-bedded calcareous lacustrine of Paleozoic oils. Cretaceous oils possibly migrated into
deposits (Cherroni, 1977; Boll, 1990). These sediments these newly formed traps, analogous to entrapment in
represent the source and hydrocarbon reservoir intervals northwestern Argentina (Bianucci et al., 1982).
in northwestern Argentina (Yacoraite and Olmedo It has been shown that transfer zones not only form
formations). In the Altiplano, the equivalent rocks of the trap door structures but also control the facies distribu-
El Molino Formation have recently been identified as a tion of reservoirs and source rocks in the depocenters
potential source rock (Edman et al., 1989; M. Cirbián, that they separate (Tankard et al., 1989; Welsink et al.,
1992, personal communication), and several reservoir 1989a). Consequently, exploration opportunities are
and seal intervals are known to exist (Figure 6). Active oil strongly affected by the transfer zones and could differ
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 315

Figure 9—Structural features


based on interpretation of
seismic lines (Figure 3) and
superimposed on the Bouguer
gravity. Note the match
between the two data sets.
A–A', B–B', C–C', and D–D' are
portions of seismic lines
shown in Figures 10, 11, 12,
and 13, respectively.

from one depocenter to the other. The negative results of trends northeast-southwest. The Uyuni-Kenyani fault
the Vilque are thus not representative of the entire zone subparallels the first-order lineaments of north-
southern Altiplano. Good exploration opportunities western Argentina, while other conspicuous transverse
remain in the area. lineaments trend northwest-southeast (Figures 2, 7, 8).
Integration of seismic, gravity, and aeromagnetic inter-
pretations of the southern Altiplano suggests that these
DISCUSSION smaller second-order transverse lineaments acted as
transfer faults during extension. On a larger scale, two
Figure 2 shows the broad-scale outline of the Creta- first-order northwest-southeast lineaments divide the
ceous rift system as it developed in the Altiplano and in Altiplano into northern, central, and southern segments
northwestern Argentina. In northwestern Argentina, (Figures 1, 2). We believe that these first-order lineaments
there is a clear relationship between tectonic lineaments, in the Altiplano functioned as first-order transfer faults to
locations of Early Cretaceous volcanism, and configura- control the style and type of basin formation (see
tion of the Cretaceous depocenters. These predominantly Tankard and Welsink, 1989).
northeast-southwest trending first-order lineaments are The analogy between the northwestern Argentinian
associated with the basement arches characteristic of the rift basins and those of the southern Altiplano is evident
Pampean terrane (Baldis, 1990). The Aconquija lineament in the seismic data. The seismic line published by
(Figure 2) separates the crystalline basement of the Bianucci et al. (1982) clearly shows the inversion of the
Pampean massif and the northwestern Argentinian rift rift sequence in the Lomas de Olmedo basin (Figure 2) by
system that developed north of it by reactivation of the reactivation of extensional faults. The rift system was
anisotropic Precambrian basement and Paleozoic rocks variably affected by the Andean deformation. In the
(Salfity, 1985). In the Altiplano, a single major lineament Lomas de Olmedo basin, inversion developed to an early

(text continues on p. 320)


316 Welsink et al.

Figure 10—Uninterpreted and interpreted seismic line 2587-28 (A–A' in Figure 9), showing a central horst flanked by half-
grabens with opposite polarities. Numbered sequences correspond to those in Figure 6 and Table 2. Line is not migrated.
(Vertical exaggeration 3:1 at 1. 5 sec two-way time.)
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 317

Figure 11—Uninterpreted and interpreted seismic line 25106-28, unmigrated (B–B' in Figure 9). Same rift geometry as in line
A–A' (Figure 10), but with incipient inversion of the basin-bounding faults. (Vertical exaggeration 3:1 at 1. 5 sec two-way
time.)
318 Welsink et al.

Figure 12—Uninterpreted and interpreted seismic line 2586-28, unmigrated (C–C' in Figure 9), showing a change in rift
geometry south of a major transverse fault between B–B' and C–C'. Note the major inversion of the rift center and the break-
through at the eastern rift margin. (Vertical exaggeration 3:1 at 1. 5 sec two-way time.)
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 319

Figure 13—Uninterpreted and interpreted seismic line 2585-28, unmigrated (D–D' in Figure 9). Widening of rift and break-
through at previously western rift margin. (Vertical exaggeration 3:1 at 1. 5 sec two-way time.)
320 Welsink et al.

Figure 14—Rift architecture


interpreted from seismic and
potential field data. Internal
transfer faults divide the basin
into three segments of
opposite polarity and variable
amounts of extension. In areas
without seismic control, the
isopachs of the Cretaceous rift
sequence have been guided
by Bouguer gravity trends.
(See Figure 9 for explanation
of symbols.)

stage and was not fully developed until it reached the 1993). Extension of the northwestern Argentinian rift
Alemania and Metán basins (Grier, 1990). This style of system into northern Chile is therefore not likely. Recent
variable deformation also characterizes the southern work and reflection seismic indicate the presence of indi-
Altiplano rift basins (Figure 15). vidual half-grabens such as the Salar de Atacama basin
The stratigraphies of the southern Altiplano and (Flint et al., 1993). The regional sag basin may have
northwestern Argentina are similar at the scale of the connected the areas prior to Andean contraction.
unconformity-bounded sequence (Figure 16). The simi- Along the length of the Andean orogen, the develop-
larity at this scale suggests that the two areas were ment of a suite of extensional basins is attributed to the
tectonically linked and shared the same processes in the Mesozoic evolution of the western continental margin of
evolution of their basins (rift and sag). Linkage was South America (Uliana et al., 1989). These basins line up
achieved through reactivation of northwest- and en echelon with rift basins across the South American
northeast-trending lineaments. However, they were not continent in the zone between the Agulhas-Malvinas
yoked together until deposition of the El Molino and transform and the Martín García high, offshore Uruguay
Yacoraite formations (Cherroni, 1977). The similarity in (Figure 17). Their location and NNW-SSE orientation
structural and stratigraphic evolution characterizes the suggest that they are related to reactivation of the
Cretaceous rift basins of southwestern Bolivia and north- Paleozoic structural fabric (Uliana et al., 1989). Although
western Argentina. This linkage suggests obvious explo- this reactivation is generally attributed to evolution of the
ration attractions of the Bolivian Altiplano. Pacific subduction margin, the influence of South
Conversely, the tectonic and stratigraphic evolution of Atlantic rifting may also have been important. Based on
the Salar de Atacama basin (Figures 2, 16) is different and the trend and ages of alkaline volcanism (148 Ma for
reflects a transition from a back-arc to a fore-arc basin Salado basin, ~128 Ma for northwestern Argentinian
setting during the Cretaceous and Tertiary (Flint et al., basin, and 82.5 Ma for Andean basin), Grier (1990) corre-
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 321

lated development of a failed rift to South Atlantic


rifting. Basalts in the Asunción basin (Figure 17) seem to
fit this trend, with an age of ~136 Ma (Wiens et al., 1993).
Figure 17 clearly shows the NNW-SSE trend of the
Triassic–Jurassic rift basins that were the first to form at
the fringes of Gondwana (Uliana et al., 1989). Northeast-
southwest transfers offset and link these basins. During
the Cretaceous, South Atlantic rifting prior to breakup
produced a northwest-trending right-lateral shear
system in which rift basins propagated inward and the
Triassic transfers were reactivated as normal faults. The
rifts are distributed as en echelon suites of basins such as
those described by Nelson et al. (1992) by the way they
interact. Following their terminology, rift jumps occur
when rifts are isolated and en echelon, such as between
the Neuquén and Cuyo basins and between the Andean
and West Peruvian basins. Rift splays can be observed
between the Cuyo and Bolsones basins, and rift offsets
are clear within the Neuquén and Colorado basins. Rift
passes occur when basin-bounding faults overlap and
encompass an intervening high. This might be the case in
the northwestern Argentinian basin, where the Salta-
Jujuy high is surrounded by rift basins. A rift gap occurs
when a connection is missing or less evident and when
the basins are collinear, such as between the Andean and
northwestern Argentinian basins.
Andean reactivation of extensional faults during
inversion resulted in a style of contractional deformation
that reflects the inheritance of the extensional fabric. In
northwestern Argentina, Grier (1990) recognized
inversion of rift-bounding faults based on stratigraphic
relationships across thrust faults and their coincidence
with basin margin trends. The association of thrust fault
vergence and rift basin margin was also noted. We
Figure 15—Schematic drawings of the seismic lines in suggest that the preference of thrust fault vergence
Figures 10, 11, 12, and 13, showing the sequential develop- caused by an extensional fabric may also be applicable to
ment of inverted structures. Numbered sequences corre- the Bolivian Eastern Cordillera. Adjacent to the
spond to those in Figure 6 and Table 2. Altiplano, a distinct westward-verging fold and thrust
belt is antithetic to the main Andean thrust system

Figure 16—Comparison of
stratigraphy, deformation
style, and timing in adjacent
areas during the Cretaceous
and Tertiary: Salar de
Atacama (northern Chile),
southern Altiplano
(southern Bolivia), and
northwestern Argentinian
basins. Arrows represent
type and timing of faulting;
white areas, shale; v,
volcanic rocks. Lithologic
symbols same as in Figure
4. (Modified after Flint et al.,
1993; Salfity and Marquillas,
1981; G. D. Vergani, 1993,
personal communication.)
322 Welsink et al.

rock deposition and for trap formation in the central and


northern Altiplano. The Paleozoic section that was
involved in the same deformational cycle may be locally
prospective.
Gravity highs are generally associated with positive
Paleozoic topography at depth. However, if the Vilque
well had indeed reached Paleozoic rocks at total depth,
then the extensive gravity low in this area would be in
contrast to the observed relationships between gravity
and basement structures (e.g., Murusa high, Figure 9).
Gravity modeling, which accounts for the presence of salt
in the San Vicente Formation, showed that the presence of
a basement high such as Murusa is unlikely. This has
several possible explanations. First, the Vilque well may
not have encountered Paleozoic rocks, suggesting the
presence of more rift sediments below total depth.
Second, a blind thrust fault could have emplaced
Paleozoic rocks. Third, the influence of topographic
elevation on gravity values increases substantially
eastward. Age dating appears to be correct, supporting
the presence of a Paleozoic thrust sheet above an untested
Cretaceous section. The exploration potential of the area
thus remains untested. A similar association may exist in
the westward-verging fold and thrust belt along the
eastern border of the Altiplano where Cretaceous rocks
are repeatedly thrusted at the surface. Within this fold
and thrust belt south of Oruro (Figure 1), the combination
of a well-developed El Molino Formation and structure at
depth may offer exploration opportunities.

Acknowledgments We would like to thank Miguel Cirbián


for his helpful suggestions and Yacimientos Petrolíferos
Fiscales Bolivianos for permission to publish this paper. The
comments of K. Biddle, P. Ziegler, C. Fernandez Garrasino, H.
Belotti, and G. Vergani are very much appreciated.
Figure 17—Composite map of Triassic and Cretaceous rift
basins showing two distinct trends as a result of different
stress regimes reactivating basement structures. (Modified REFERENCES CITED
after Cherroni, 1977; Salfity, 1985; Macellari, 1988; Uliana et
al., 1989.)
Allen, P. A., P. Homewood, and G. D. Williams, 1986,
Foreland basins: an introduction, in P. A. Allen and P.
Homewood, eds., Foreland basins: International Associa-
(Roeder, 1988) and contains a well-developed Cretaceous tion of Sedimentologists Special Publication 8, p. 3–12.
rift and sag sequence. This suggests that eastward- Allmendinger, R. W., V. A. Ramos, T. E. Jordan, M. Palma,
dipping extensional faults of a western rift margin were and B. L. Isacks, 1983, Paleogeography and Andean struc-
inverted. The eastward change to eastward-verging tural geometry, northwest Argentina: Tectonics, v. 2,
thrust faults might reflect the change to westward- p. 1–16.
dipping extensional faults of the opposite basin margin. Aranibar, O., and E. Martinez, 1990, Structural interpretation
Structural inversion involves the reactivation of a of the Altiplano, Bolivia (abs. ), in Structure and evolution
preexisting extensional fault system during compression. of the Atlas Mountain system in Morocco and structure
The presence of Cretaceous rift sediments in the main and evolution of the central Andes in northern Chile and
structures of the Altiplano, such as Uyuni-Kenyani, southern Bolivia and northwestern Argentina: Final
workshop of the research group on mobility of active
Sevaruyo, and Andamarca (Figure 1), suggests that these continental margins, Freie Universität Berlin, Germany,
structures are the result of inversion, and similar struc- p. 47.
tures are believed to be present in the subsurface of the Baldis, B. A., 1990, Relaciones de estructuras de cizalla
central and northern Altiplano. The development of the pampeanas con la placa sudamericana y el continente
extensional architecture and its subsequent inversion Gondwana: Décimo Primer Congreso Geológico
created optimal conditions for source rock and reservoir Argentino, San Juan, Argentina, v. 2, p. 301–305.
Structural Inversion of a Cretaceous Rift Basin, Southern Altiplano, Bolivia 323

Bally, A. W., 1982, Musings over sedimentary basin evolution: Roeder, D., 1988, Andean-age structure of Eastern Cordillera
Royal Society of London Philosophical Transactions, Series (Province of La Paz, Bolivia): Tectonics, v. 7, p. 23–29.
A, v. 305, p. 325–338. Salfity, J. A., 1985, Lineamientos transversales al rumbo
Bianucci, H., J. F. Homovc, and O. M. Acevedo, 1982, andino en el noroeste argentino: Cuarto Congreso
Inversión tectónica y plegamientos resultantes en la Geológico Chileno, Antofagasta, Chile, v. 2, p. 119–137.
comarca Puesto Guardián–Dos Puntitas, departamento Salfity, J. A., and R. A. Marquillas, 1981, Las unidades estrati-
Orán, Provincia de Salta: Primer Congreso Nacional de gráficas cretácicas del norte de la Argentina, in W.
Hidrocarburos, Petróleo y Gas, Exploración, Buenos Aires, Volkheimer and E. A. Musacchio, eds., Cuencas sedimen-
Argentina, p. 23–30. tarias del Jurásico y Cretácico de América del Sur: Comité
Boll, A., 1990, Identificación y correlación de secuencias Sudamericano del Jurásico y Cretácico, Buenos Aires,
somerizantes Miembro Las Avispas (Formación Yacoraite) Argentina, v. 1, p. 303–317.
noroeste argentino: Décimo Primer Congreso Geológico Scott, D. L., and B. R. Rosendahl, 1989, North Viking graben:
Argentino, San Juan, Argentina, v. 2, p. 153–156. an east African perspective: AAPG Bulletin, v. 73,
Cady, J. W., and R. A. Wise, 1992, Gravity and magnetic p. 155–165.
studies, in Geology and mineral resources of the Altiplano Sheffels, B. M., 1988, Structural constraints on crustal short-
and Cordillera Occidental, Bolivia: USGS Bulletin 1975, ening in the Bolivian Andes: Ph. D. dissertation, Massa-
p. 56–62. chusetts Institute of Technology, Cambridge, MA, 170 p.
Cherroni, C., 1977, El sistema cretácico en la parte boliviana Tankard, A. J., and H. J. Welsink, 1987, Extensional tectonics
de la cuenca cretácica andina: Revista Técnica de and stratigraphy of Hibernia oil field, Grand Banks,
Yacimientos Petrolíferos Fiscales Bolivianos, La Paz, Newfoundland: AAPG Bulletin, v. 71, p. 1210–1232.
Bolivia, v. 6, p. 5–46. Tankard, A. J., and H. J. Welsink, 1989, Mesozoic extension
Edman, J. D., J. R. Kirkpatrick, D. D. Lindsey, J. D. Lowell, M. and styles of basin formation in Atlantic Canada, in A. J.
Cirbián, and M. Lopez, 1989, Hydrocarbon potential of Tankard and H. R. Balkwill, eds., Extensional tectonics and
Altiplano and northern Subandean, Bolivia (abs. ): AAPG stratigraphy of the north Atlantic margins: AAPG Memoir
Bulletin, v. 73, p. 352. 46, p. 175–195.
Flint, S., P. Turner, E. J. Jolley, and A. J. Hartley, 1993, Exten- Tankard, A. J., H. J. Welsink, and W. A. M. Jenkins, 1989,
sional tectonics in convergent margin basins: an example Structural styles and stratigraphy of the Jeanne d’Arc
from the Salar de Atacama, Chilean Andes: GSA Bulletin, basin, Grand Banks of Newfoundland, in A. J. Tankard
v. 105, p. 603–617. and H. R. Balkwill, eds., Extensional tectonics and strati-
Gibbs, A. D., 1984, Structural evolution of extensional basin graphy of the north Atlantic margins: AAPG Memoir 46,
margins: Journal of Geologic Society of London, v. 141, p. 265–282.
p. 609–620. Thorpe, R. S., P. W. Francis, and R. S. Hamon, 1981, Andean
Grier, M. E., 1990, The influence of the Cretaceous Salta rift andesites and crustal growth: Royal Society of London
basin on the development of Andean structural geome- Philosophical Transactions, Series A, v. 301, p. 305–320.
tries, NW Argentine Andes: Ph.D. dissertation, Cornell Uliana, M. A., K. T. Biddle, and J. Cerdán, 1989, Mesozoic
University, Ithaca, NY, 178 p. extension and the formation of Argentine sedimentary
Isacks, B. L., 1988, Uplift of the central Andean plateau and basins, in A. J. Tankard and H. R. Balkwill, eds., Exten-
bending of the Bolivian orocline: Journal of Geophysical sional tectonics and stratigraphy of the north Atlantic
Research, v. 93, p. 3211–3231. margins: AAPG Memoir 46, p. 599–614.
Macellari, C. E., 1988, Cretaceous paleography and the deposi- Welsink, H. J., J. D. Dwyer, and R. J. Knight, 1989a, Tectono-
tional cycles of western South America: Journal of South stratigraphy of the passive margin off Nova Scotia, in A. J.
American Earth Sciences, v. 1, p. 373–418. Tankard and H. R. Balkwill, eds., Extensional tectonics and
Marquillas, R. A., and J. A. Salfity, 1988, Tectonic framework stratigraphy of the north Atlantic margins: AAPG Memoir
and correlations of the Cretaceous–Eocene Salta Group; 46, p. 215–231.
Argentina, in H. Bahlburg, Ch. Breitkreuz, and P. Giese, Welsink, H. J., S. P. Srivastava, and A. J. Tankard, 1989b, Basin
eds., The southern central Andes: Lecture Notes in Earth architecture of the Newfoundland continental margin and
Sciences 17, Heidelberg, Germany, Springer-Verlag, its relationship to ocean crust fabric during extension, in A.
p. 119–136. J. Tankard and H. R. Balkwill, eds., Extensional tectonics
Moreno, J. A., 1970, Estratigrafía y paleogeografía del and stratigraphy of the north Atlantic margins: AAPG
Cretácico Superior en la cuenca del noroeste argentino, con Memoir 46, p. 197–213.
especial mención a los subgrupos Balbuena y Santa Wiens, F., M. E. González, and R. Ruff, 1993, Desarrollo
Bárbara: Revista Asociación Geológica Argentina, Buenos tectono-sedimentario del bloque de Asunción, Paraguay:
Aires, Argentina, v. 25, p. 9–44. Décimo Segundo Congreso Geológico Argentina y
Nelson, R. A., T. L. Patton, and C. K. Morley, 1992, Rift- Segundo Congreso de Exploración de Hidrocarburos,
segment interaction and its relation to hydrocarbon explo- Mendoza, Argentina, v. 1, p. 27–32.
ration in continental rift systems: AAPG Bulletin, v. 76, Williams, G. D., C. M. Powell, and M. A. Cooper, 1989,
p. 1153–1169. Geometry and kinematics of inversion tectonics, in M. A.
Reyes, F. C., 1972, Correlaciones en el Cretácico de la cuenca Cooper and G. D. Williams, eds., Inversion tectonics:
andina de Bolivia, Perú y Chile: Revista Técnica de Geological Society Special Publication 44, p. 3–15.
Yacimientos Petrolíferos Fiscales Bolivianos, La Paz, Yalamanchili, S. V., E. Martinez, and O. Aranibar, 1991, Aero-
Bolivia, v. 1, p. 101–144. magnetic structural interpretation and evaluation of
Reyes, F. C., and J. A. Salfity, 1973, Consideraciones sobre la hydrocarbon and mineral prospects, Altiplano, Bolivia
estratigrafía del Cretácico (subgrupo Pirgua) del noroeste (abs. ): Society of Exploration Geophysicists 61st Annual
argentino: Quinto Congreso Geológico Argentino, Buenos International Meeting and Exposition, Houston, SEG
Aires, Argentina, v. 3, p. 355–385. Abstracts 61, p. 633–635.
324 Welsink et al.

Authors’ Mailing Addresses


H. J. Welsink
Perez Companc
C. C. 181
8300 Neuquén
Argentina

E. Martinez
O. Aranibar
J. Jarandilla
Yacimientos Petrolíferos Fiscales Bolivianos
Casilla 1659
Santa Cruz de la Sierra
Bolivia
Geometry and Seismic Expression of the Cretaceous
Salta Rift System, Northwestern Argentina

Alberto H. Comínguez Victor A. Ramos


CONICET Universidad de Buenos Aires
Universidad Nacional de La Plata Buenos Aires, Argentina
La Plata, Argentina

Abstract

T he foothills of the central Andes of northwestern Argentina hinder the interpretation of the complex
structural rift system developed during late Mesozoic extension. Andean compressive deformation
inverted the Salta rift system, resulting in a series of complex structures with trends oblique to the main Andes.
The Lomas de Olmedo basin, a failed branch of the rift system located east of the Andean orogenic front, was
selected to undertake deep reprocessing of the available industrial seismic lines. A 150-km-long seismic section
of the basin, recorded with Vibroseis and dynamite sources, was reprocessed. Extended correlation applied to
the Vibroseis seismic data yielded reliable results down to 9 sec two-way travel time. Acoustic horizons identi-
fied within this interval include the deepest synrift deposits in the axial part of the basin and a deep oblique
discontinuity in the crust. On this basis, a complete cross section of the basin was made. This study documents
the asymmetry of the rift, with a prominent zone of thermal uplift in the northern edge. Truncation of the
Paleozoic beds and identification of a deep oblique discontinuity at 7–8 sec (18–21 km deep) suggest that a
northward-dipping detachment controlled the asymmetry of the system. The rift structure is mildly modified
by folding related to Cenozoic tectonic inversion in the southern sector of the basin. This inversion was
controlled mainly by strike-slip displacements along the previous normal faults.

Resumen

E l pie de monte de los Andes Centrales del noroeste de Argentina esconde la interpretación de un
complejo sistema estructural desarrollado durante la extensión mesozoica tardía. La deformación
compresiva ándica invirtió el sistema del rift Salta, produciendo una serie de estructuras complejas con arrum-
bamientos oblícuos a los Andes. La cuenca de Lomas de Olmedo, una rama abortada del sistema de rift local-
izado al este del frente orogénico ándico, fue seleccionada para realizar un reprocesamiento profundo de lineas
sísmicas disponibles de la industria. En particular fue seleccionada una línea sísmica de la cuenca de 150 km de
longitud, registrada en su mayor parte con Vibroseis y complementada con dinamita. Un algoritmo de
correlación extendida se aplicó a los datos sísmicos de Vibroseis alcanzándose resultados confiables hasta
profundidades correspondientes a 9 segundos de tiempo de travesía doble. Los horizontes acústicos identifi-
cados dentro de este intervalo de tiempo incluyeron tanto los depósitos más profundos de sinrift en la parte
axial de la cuenca, como una profunda discontinuidad oblícua en la corteza. Sobre esta base fue confeccionada
una sección completa de la cuenca. Los resultados de este estudio documentaron la asimetría del rift, con una
zona prominente de levantamiento térmico a lo largo del borde norte. El truncamiento de los estratos paleo-
zoicos, así como la identificación de una profunda discontinuidad oblícua a los 7–8 segundos de doble tiempo
de travesía (aproximadamente 18–21 km), sugieren que un nivel de despegue inclinado hacia el norte controló
la asimetría del sistema. La estructura de rift fue suavemente modificada por el plegamiento relacionado a la
inversión tectónica cenozoica en el sector sur de la cuenca. Esta inversión fue en parte acomodada preferente-
mente por desplazamientos de rumbo a lo largo de las fallas normales previas.

Comínguez, A. H., and V. A. Ramos, 1995, Geometry and seismic expression of the Creta- 325
ceous Salta rift system, northwestern Argentina, in A. J. Tankard, R. Suárez S., and H. J.
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 325–340.
326 Comínguez and Ramos

INTRODUCTION
The Salta rift comprises a complex suite of extensional
troughs that characterize the Late Cretaceous–early
Tertiary paleogeography of northwestern Argentina. The
western part of the rift was inverted during the Andean
deformation in the late Tertiary and has since been
involved in the complex fold and thrust belt. To under-
stand the original structure of the Salta rift system, it was
necessary to study the easternmost branch, adjacent to
the present orogenic front of the Andes. The purpose of
this study is to document the geometry and seismic
expression of the Lomas de Olmedo basin in an area
where compressional deformation was only mild or
nonexistent (Figure 1). We have used these data to
interpret the tectonic setting and evolution of the rift
system.
This research is part of the CONICET–CAPLI
(Argentine National Science Foundation, Argentine
Committee for the Lithosphere) program undertaken as
a join venture by Yacimientos Petrolíferos Fiscales
(YPF S.A.) and the Universities of La Plata and Buenos
Aires. The aims of this research program are to study
and reprocess seismic reflection data that have a poten-
tially deep record.

Figure 1—Isopach map of the Salta rift system and its rela-
tionship to the Andean basin of Bolivia. The outlines of the
FIELD WORK AND PROCESSING Amazonia and Arequipa cratons are shown. (Based on
A 150-km-long section of seismic data from three field Salfity, 1982.)
programs acquired by YPF S.A. from 1981 to 1983 were
reprocessed (Figures 2, 3). Vibrators were used as a
source in the southern sector of the line; the remaining
record involved dynamite. Final collection of data from
these programs required major efforts from both the
mathematical processing of multichannel seismic infor-
mation and monitoring of the preliminary results prior to
the final seismic section (Figure 4).
The YPF Alto de la Sierra well (ADSL x-1) was used as
a control to constrain interpretation of the various
acoustic horizons. Geologic information from several
abandoned wells near the line contributed to a broader
understanding of the seismic data. In the Appendix we
desribe the mathematical reprocessing, its relevance, its
limitations on interpretation, and its impact on the inter-
pretation of the Cretaceous rift.

TECTONIC FRAMEWORK
The location of the Cretaceous Andean basins and the
Salta rift system in this part of South America is
controlled by late Precambrian–Cambrian tectonic
fabrics. This is expressed in the various Precambrian
cratons that were amalgamated during the Pampean
orogeny in the Early Cambrian. The complex distribu-
tion of the basins is partially controlled by a triple Figure 2—Isopach map of Lomas de Olmedo basin and
junction involving the Amazonia, Arequipa, and Pampia Salta rift system showing the Andean orogenic front. West
cratons (Ramos and Vujovich, 1994). The Lomas de of the front is a severe inversion of the rift. Also shown is
Olmedo basin subsided along the suture between the the location of the seismic line. (Based on Salfity, 1980;
Amazonia and Pampia cratons. Gómez Omil et al., 1989; Ferreiro, 1989; Carle et al., 1991.)
Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 327

← Figure 3—Prerift geologic map showing location of


seismic lines 04196 and 02128. This later line was also
reprocessed and contributed to the general interpretation
of the basin. (Modified after Carle et al., 1991.)

Several publications have discussed the Salta rift


system, but the tectonic setting of the Lomas de Olmedo
basin has only been adressed by Bianucci et al. (1981).
These authors have attributed the geometry of the Salta
rift to hotspots that controlled the inception of the
different branches. Traditionally, these basins have been
described as an intricate combination of topographic
highs and lows. This paleogeography has been docu-
mented in detail by Moreno (1970), Salfity (1979, 1980,
1982), Salfity et al. (1985), and Salfity and Marquillas
(1986). However, as noted by Allmendinger et al. (1982),
these paleogeographies have not been palinspastically
restored to accommodate Andean shortening. Only
recently have Grier and Allmendinger (1991) attempted
to restore a complete cross section of the basin. Neverthe-
less, their restoration does not include the west-east
asymmetry and the oblique nature of the master detach-
ment of the basin.

(below)
Figure 4—Migrated seismic line 4196 with vertical exagger-
ation of about 4× to enhance the relationship between
prerift and synrift deposits. Note that in the axial part of the
basin (CDP 4200–4500), the seismic character of the synrift
deposits contrast with the basement fabric. See geologic
interpretation in Figure 7.
328 Comínguez and Ramos

Table 1—Subsurface Paleozoic Unitsa

Northwestern Sector Age Southeastern Sector


Huamampampa
Formation Devonian Tonono Formation
Icla Formation Devonian Rincón Formation
Santa Rosa Formation Devonian Caburé Formation
Kirusillas Formation Silurian Copo Formation
Las Breñas Formation Cambrian– Las Breñas
Ordovician Formation
aAfter Mingramm et al. (1979).

Due to the substantial Andean shortening of the basin


west of the orogenic front (Figure 3), it is difficult to
establish the original structure of the basin. Conse-
quently, special attention was given to the Lomas de
Olmedo basin, a branch of the rift located east of the
present Andean thrust front.

SALTA RIFT STRATIGRAPHY Figure 5—Strength attribute by color demodulation


analysis of a window in the northern sector of line 4196
The Salta rift basins have a simple stratigraphy that showing basement seismic character of Precambrian
has been known since the early work of Moreno (1970), basement. (A) Coherency of the Ordovician top reflectors
Reyes and Salfity (1973), and Reyes et al. (1976). The represent the strongest and most persistent acoustic
stratigraphy involved in the deformation is divided in horizon in the study area. (B) An angular unconformity
exists between Paleozoic prerift and (C) Cretaceous rift
this paper into prerift, synrift, and postrift sequences.
deposits (see interpretation in Figure 6).

Prerift Sequences
the east of the study area. The Silurian and Devonian
The western part of the rift succession was deposited sequences were deposited in a foreland basin east of the
over the Puncoviscana Formation. This unit comprises Ocloyic orogenic front. This front developed during the
Precambrian–Lower Cambrian turbiditic deposits that Late Ordovician collisional event, which was related to
fringe the western edge of the middle–upper Precam- final docking of the Arequipa terrane (Ramos, 1986).
brian Pampia craton. The Lomas de Olmedo basin was There is a strong contrast in strength attribute between
developed on lower Paleozoic deposits that were previ- the Precambrian basement character and the lateral
ously deformed during the Chanic orogeny at the end of coherency of the Cambrian–Ordovician deposits, as
the Devonian (Turner and Méndez, 1975). determined by color demodulation analysis in the north-
The subsurface map traversed by the seismic lines western area (Figure 5). The seismic velocity of these
shows that the present trend of the subcrops reflects the rocks ranges between 5800 and 6000 m/sec. It is apparent
unconformity between the synrift and the Paleozoic that the basement consists of crystalline Precambrian
deposits (Figure 4). Although this unconformity and the rocks of the Pampia craton, similar to that intersected in
present outcrop pattern were well known in the sub- several exploration wells farther east (Russo et al., 1979).
Andean region, as recognized by Reyes (1959) in The seismic characteristics of the various Paleozoic
northern Argentina and southern Bolivia, it was tradi- units are shown in Figure 6. One of the most conspicuous
tionally attributed to a pre-Cretaceous folding since acoustic horizons corresponds to the top of the Las
Padula’s (1959) work. However, the approximate east- Breñas Formation (Table 1). The lateral coherence traces
west orientation does not fit any of the Paleozoic trends this horizon throughout most of the rift system
of deformation mapped in northwestern Argentina by (Acevedo, 1986). Seismic interval velocities for these
various authors (e.g., Turner and Méndez, 1975). deposits computed from stacking velocity analyses range
Bianucci and Homovc (1982) were the first to recognize from 4100–4500 m/sec in the northwestern part of the
that the prerift doming, uplift, and differential erosion basin to more than 4600 m/sec in the central and deeper
were a consequence of an early stage of Cretaceous rift parts of the seismic line.
development. The southeastern part shows that subhorizontal beds
In the present study area and along the seismic lines, of the Las Breñas Formation are preserved with a thin
the units listed in Table 1 are recognized. This strati- cover of Copo Formation at the southern end of the line
graphy is combined with the succession identified in the (south of CDP 5.400). A strong angular unconformity
sub-Andean belt of Argentina and Bolivia. The Cambrian– truncates the Paleozoic deposits in the northern area
Ordovician deposits belong to a westward-facing clastic (Figures 5, 6), while in the southern area it is represented
platform, with the shore zone located a few kilometers to by a paraconformity (Figure 7).
Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 329

Figure 6—Uninterpreted and interpreted migrated seismic line 4196, northwestern part, with vertical exaggeration of about
4× to enhance the Paleozoic stratigraphy. Vertical scale is two-way time, in seconds.
330 Comínguez and Ramos

Figure 7—(Top) Migrated seismic line 4196; see Figure 4 for location. (Bottom) Geologic interpretation of the seismic line and
location of the Alto de La Sierra x-1 exploration well. (continues across facing page)

Synrift Deposits Aconquija lineaments. This Early Cretaceous volcanism


constitutes the first volcanic cycle of Reyes et al. (1976)
The synrift deposits are represented by different units and Salfity (1982) (Figure 8). Volcanic rocks crop out in
of the Pirgua Subgroup. This subgroup is composed of the Alemanía and Metán basins where they are repre-
proximal fanglomerates, fluvial conglomerates, and red sented by two different pulses in the El Cadillal and La
sandstones and shales associated with alkaline basaltic Yesera formations. The first is characterized by alkaline
rocks (Figure 8). trachytes, in part perpotassic, which have ages of
112–128 Ma, while the second is composed of K-foidites
Volcanic Rocks (nepheline-leucite normative) of 97–103 Ma according to
The chemistry of the volcanic rocks has been studied Galliski and Viramonte (1985, 1988). Farther north along
in detail by Galliski and Viramonte (1988). These rocks the topographic high between Lomas de Olmedo and El
represent within-plate volcanism associated with major Rey basins, the Cachipunco basalt is exposed.
lineaments that bound the rift, such as the Isonza and This first magmatic cycle, contemporaneous with sedi-
Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 331

Figure 7 (continued)

mentation of the Pirgua Subgroup, is also associated with consists of lava flows of trachyandesitic composition
alkaline granitoids (the 96–155 Ma El Aguilar, Cobres, intercalated with the Los Blanquitos Formation of the
and Tusaquillas stocks; Méndez et al., 1979). The ages of Tres Cruces basin (Coira, 1979).
basaltic volcanism decrease toward the east (Bossi and A third volcanic cycle is widespread at Río Capillas
Wampler, 1969; Reyes et al., 1976; Valencio et al., 1976). over the Salto-Jujeña high (Figure 2) along the Puntitas–
The second magmatic cycle of the Salta Group (Salfity, El Totoral lineament of the Lomas de Olmedo basin
1982) is also contemporaneous with the Pirgua Subgroup (Bianucci et al., 1981) and the Alemanía basin (63–55 Ma)
and has two eruptive centers. One is located in the (Cortelezzi et al., 1979; Omarini et al., 1987). The rocks of
depocenter of the Alemanía basin (see Figure 3) and is the third volcanic cycle are mainly basalts (Cortelezzi et
represented by the Las Conchas basalt (76–78 Ma); it also al., 1979) and lamproitic dikes (Omarini et al., 1987). A
interfingers with the upper levels of the Las Curtiembres fourth volcanic cycle, consisting of volcanic rocks
Formation. The Las Conchas basalt is volumetrically the emplaced in the Yacoraite Formation, is represented by
more important unit. It is characterized by basanites, the Palmar Largo volcanics. These volcanics were
mugearites, and tephrifonolites with abundant peridotite emplaced during reactivation of normal faulting and are
xenoliths. These extrusives were explosive and had composed of olivine basalts, andesites, and leucoan-
multiple centers of eruption. The other eruptive center delacites (Carle et al., 1991).
332 Comínguez and Ramos

The evolution of these volcanic rocks records Table 2—Interval Velocity and Depth Versus Timea
migration of the volcanic centers from the northeast-
trending fault margins of the various basins in the Early Two-Way Time Interval Velocity Depth
Cretaceous (first volcanic cycle), to the axis of maximum (msec) (m/sec) (m)
subsidence in the Alemanía and Tres Cruces basins by 0–825 2081 0–858
the end of the Cretaceous (second volcanic cycle). The 825–1200 2630 858–1351
third and fourth volcanic cycles migrated to the eastern- 1200–1425 2846 1351–1671
most basin and were restricted to the edge of the Salto- 1425–2475 3568 1671–3544
Jujeña high and the axis of the Lomas de Olmedo basin. 2475–2925 3979 3544–4439
The general trend within the Pirgua Subgroup is from 2925–3150 4112 4439–4901
peralkaline to subalkaline rocks, indicating greater 3150–3450 4576 4901–5587
melting in the source associated with increasing 3450–5700 5767 5587–12074
extension and larger eruption volumes. The fourth cycle 5700–7350 ~7000 12074–~18000
is again more alkaline, showing minor reactivation of the
aLINE 04196, CDP = 4385.
rifting.

Sedimentary Rocks of isotopic dating of the interbedded volcanics, a


In the Alemania outcrop area, a 50˚–74˚ angular Valanginian–Santonian age is assigned to the Pirgua
unconformity separates the base of the Pirgua Subgroup Subgroup.
from the prerift basement (Grier et al., 1987). Seismic
velocities in the lower Pirgua Subgroup in the Lomas de Postrift Deposits
Olmedo basin range from 3.3 to 4.0 m/sec and may
exceed 4.5 m/sec in the deeper parts of the basin. Postrift deposits are represented by the Balbuena and
The various lithologic units of the Pirgua Subgroup Santa Bárbara subgroups. The Balbuena Subgroup repre-
shown in Figure 8 consist of two major tectonosedimen- sents an early sag phase, partially interrupted by minor
tary sequences (Gómez Omil et al., (1989). A lower fault reactivation, that was also responsible for the
sequence (La Yesera and Las Curtiembres formations) Palmar Largo basalts. The Balbuena Subgroup is
corresponds to alluvial fan facies along the active composed of the Lecho, Yacoraite, and Olmedo forma-
margins of the rift, but laterally interfingers with fluvial, tions. Some authors (e.g., Gómez Omil et al., 1989),
eolian, and distal playa facies in the central part of the separate the Olmedo Formation from this subgroup due
basin. The first unit was deposited over a surface with to the unconformity that intervenes between the Olmedo
considerable topographic relief. The second unit (Los and underlying units.
Blanquitos and related formations) has similar but lower The Balbuena Subgroup consists of carbonates and
energy facies and fine-grained deposits that accumulated subaqueous clastic facies that onlap a major unconfor-
on a subtler relief due to the filling of the basin and to mity above the Pirgua Subgroup and older Paleozoic
less active tectonics. Both units together have a rocks. The areal distribution of the Balbuena Subgroup
maximum thickness greater than 5000 m as inferred from greatly oversteps the original margins of the basin
the interval velocities at CDP 4385 (Table 2). On the basis (Figure 8). The various formations are characterized by

Figure 8—Tectonostratigraphy of the Salta rift.


Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 333

tabular geometries and lateral persistence. The Yacoraite Subgroup. However, in the southern half of the area,
Formation represents a shallowing-upward sequence of there is evidence of faulting in the Balbuena Subgroup.
carbonate and subaqueous clastic facies. Eolian clastics When analyzed in detail, it is clear that the faults are
typify the Lecho Formation (Gómez Omil et al., 1989). mostly associated with minor folding (Figure 10)
The Olmedo Formation is composed of black shales and (Chiarenza and Ponzoni, 1989). These folds are attributed
anhydritic evaporites that reach a maximum thickness in to contraction related to strike-slip displacement along
the central part of the basin. This formation developed in the previous normal faults. Incipient positive flower
hypersaline lakes in the depocenters of the basin which structures are locally associated with these faults. Strike-
were surrounded by mudflats, while the northern slip displacements are related to interaction of Andean
margin of Lomas de Olmedo basin received some fluvial compression and the oblique west-southwestward trend
sands. Based on a rare microfauna, a Campanian–Maas- of the rift structures.
trichtian age is inferred for the Yacoraite Formation. Beneath these basin-forming faults along the southern
The upper sag deposits are represented by several part of the line, there is a conspicuous subhorizontal
units of the Santa Bárbara Subgroup. The Mealla, Maíz reflector between 7 and 8 sec (about 18–21 km depth)
Gordo, and Lumbrera formations record the evolution of (Figure 11). This acoustic marker separates relatively
a closed basin (Cazau et al., 1976). Their shales, marls, transparent (except for faults) upper crust from highly
and local carbonates were deposited in a suite of reflective lower crust (stacking velocities show an
ephemeral lakes with variable discharge and flooding increase in interval velocity that coincides with the lower
that alternated with fluvial deposits. Based on their fossil zone). Similar reflection patterns elsewhere have also
fauna, a Paleocene–Eocene age is assigned to the Santa been interpreted as the interface between upper and
Bárbara Subgroup. The sag phase succession is more lower crust (e.g., Peddy and Keen, 1987). For example,
than 2200 m thick. compare the intracrustal boundary shown in Figure 11
The basin was finally filled by a tabular body of distal with that illustrated by Hall (1989, his figure 1).
fluvial sediments of late Cenozoic age characterized by The top of the reflective zone dips northward at 3.5˚
seismic interval velocities ranging from 2600 to 2800 beneath the depocenter. This dipping reflector is inter-
m/sec. These are the synorogenic deposits related to preted as parallel to the detachment zone that controls
successive uplift and east-shifting of the Andean the rift system. However, it is not visible in the northern
orogenic front. This front is presently located more than part of the reprocessed seismic line because of insuffi-
120 km west of the seismic line. cient source energy and possibly because, with increasing
depth, the discontinuity loses its identity and is dis-
persed in a ductile lower crust.
STRUCTURE OF LOMAS DE OLMEDO
BASIN
TECTONIC EVOLUTION OF THE
An outstanding feature of the study area is the SALTA RIFT
asymmetry of the rift system (Figure 7). This asymmetry
is evidenced by the degree of doming preserved in the The seismic evidence shows that northwestern
prerift deposits, as well as in the distribution and vertical Argentina was subjected to extension during most of
throws of the normal faults. Doming was greater along Cretaceous and early Tertiary time (see Uliana et al.,
the northern margin of the basin (Figures 5, 6), as shown 1989). Compared with neighboring rift systems, such as
by the present truncation of the Paleozoic units. In the Chaco-Paraná rift of northeastern Argentina, north-
contrast, truncation is not recognized in the southern western Argentine rifting may have been linked in time
part of the section (Figure 9), suggesting that doming and space to the extensional regime responsible for the
was mild or nonexistent. opening of the South Atlantic Ocean. Extensional
This truncation is noticeable not only in the subsurface stretching of these basins decreased at these latitudes
of the basin but has also been mapped in the Sierras toward the west. This is inferred from the magnitude of
Subandinas, west of Tartagal (Reyes, 1959, 1978). This the β-factor, the chemical composition of the basaltic
topography has been recognized as the Michicola arch lavas (alkaline basalts in the Salta system and continental
(Vilela, 1967), but its origin is controversial. Salfity (1980) flood basalts of tholeiitic composition in the Chaco-
attributed this arch to the Late Ordovician orogenic front, Paraná basin), and the relative volume of the basalts,
while other authors have interpreted it as a peripheral with a minimum in the Salta rift.
bulge associated with Andean shortening. The west- The asymmetry of the rift is partially controlled by the
northwest trend of the Michicola arch is parallel to the west-northwest trend of the suture between the
rift basin, and its age is older than the rift deposits. Its Amazonia and Pampia cratons. This suture has the same
regional distribution indicates a genetic link with orientation as the Michicola arch. Ramos and Vujovich
doming prior to the tectonic collapse of the basin. (1994) argue that the late Precambrian–Early Cambrian
A series of east-west trending normal faults bound suture between these two blocks is northward dipping,
both flanks of the basin. The northern faults have greater implying that the main detachment of the rift system
throws than their southern counterparts. Most of the developed synthetically to the earlier subduction zone.
faults cease normal displacement by the top of the Pirgua The focus of doming at the northern margin of the
Figure 9—Uninterpreted and interpreted migrated seismic line 4196, southeastern part, with vertical exaggeration of about 4×
to enhance the Paleozoic stratigraphy. Compare the structural attitude of the subhorizontal Paleozoic deposits with the tilted
attitude shown in Figure 6.
Figure 10—Uninterpreted and interpreted migrated seismic line 4196, central sector, with vertical exaggeration of about 4×.
to show the incipient tectonic inversion of the rift system. Note the minor folding near the normal faults, indicating possible
contraction related to strike-slip displacements.
336 Comínguez and Ramos

Figure 11—Window of the deep southern sector of the seismic line of Figure 2, showing the inclined seismic discontinuity
between 7 and 8 sec two-way time.

basin establishes the area of maximum thermal uplift An understanding of the thermal history of the basin
and highest thermal gradients prior to the collapse of the and the consequent rift and synorogenic filling indicates
rift (Figure 12). This normally coincides with the area of that the northern part of the Lomas de Olmedo basin had
maximum lithospheric attenuation (Wernicke and Tilke, maximum lithospheric thinning, with relatively large
1989; Kusznir et al., 1987). As the detachment theory additions of heat into crust. This asymmetry must be
predicts, however, maximum lithospheric attenuation taken into account when evaluating the hydrocarbon
may be offset from the locus of fault-controlled subsi- potential of Lomas de Olmedo basin.
dence. In the Lomas de Olmedo rift system, the litho-
spheric attenuation is north of the fault-controlled axis of
the basin. The position of the doming is consistent with
the northward-dipping, oblique, deep reflector that Acknowledgments The authors wish to express their
presently controls the asymmetry of the rift. gratitude to the management of YPF S.A. for their logistical
The synrift deposits are more than 5000 m thick above support of the CAPLI research program, as well as to Cristina
the rift axis. The fault-controlled subsidence that accom- Vistalli, Hugo A. Bianucci, Osvaldo Acevedo, and Jorge
modated this synrift fill is also associated with alkaline González Naya for their helpful comments during the research.
basaltic volcanism (Fraga and Introcaso, 1990). Although The authors acknowledge Teresa Jordan, Suzanne M. Kay, and
sedimentation of the Balbuena Subgroup marked the Andrés Boll for their review of the manuscript. This project
beginning of the sag phase due to thermal subsidence, was also supported by the Antorchas Foundation of Argentina.
interfingering of alkaline basalts at Palmar Largo with
Yacoraite limestones indicates temporary reactivation of
the tensional regime. The basalts are mainly associated REFERENCES CITED
with normal faults at the southern margin of the rifting.
The Santa Barbara Subgroup marks the closing of the sag Acevedo, O. M., 1986, El precarbónico en la Provincia de
phase when the basin was completed filled. Salta: Boletín de Informaciones Petroleras, Tercera Serie, v.
The area received more than 2700 m of distal synoro- 6, p. 65–72.
genic deposits during late Cenozoic Andean deformation Allmendinger, R., T. Jordan, M. Palma, and V. A. Ramos,
and encroachment of the sub-Andean thrust front. 1982, Perfil estructural en la Puna Catamarqueña (25–27˚S),
Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 337

(a)

(b)

Figure 12—Schematic section of the Lomas de Olmedo rift. (a) Reconstruction of doming prior to the collapse of the rift.
High thermal gradient was related to lithospheric attenuation. (b) Structure of the rift prior to Andean tectonic inversion.

Argentina: Quinto Congreso Latinoamericano de Geología, oriental de las provincias de Salta y Jujuy: SextoCongreso
Actas I, p. 499–518. Geológico Argentino, Actas I, p. 341–356.
Bianucci, H. A., and J. F. Homovc, 1982, Tectonogénesis de un Chiarenza, D. G., and E. Ponzoni, 1989, Contribución al
sector de la cuenca del Subgrupo Pirgua, noroeste conocimiento geológico de la Cuenca Cretácica en el
argentino: Quinto Congreso Latinoamericano de Geología, ámbito oriental de la Subcuenca de Olmedo, provincia de
Actas I, p. 539–545. Salta, Republica Argentina: Primer Congreso Nacional de
Bianucci, H. A., O. M. Acevedo, and J. J. Cerdan, 1981, Exploración de Hidrocarburos, Actas I, p. 209–228.
Evolución tectosedimentaria del Grupo Salta en la Claerbout, J. F., 1976, Fundamentals of Geophysical Data
Subcuenca de Lomas de Olmedo (Provincias de Salta y Processing: with Applications to Petroleum Prospecting:
Formosa): Octavo Congreso Geológico Argentino, Actas New York, McGraw Hill, 187 p.
III, p. 159–172. Coira, B. L., 1979, Descripción geológica de la Hoja 3c Abra
Bossi, G. E., and M. Wampler, 1969, Edad del Complejo Alto Pampa, provincia de Jujuy: Servicio Geológico Nacional,
de Las Salinas y Formación El Cadillal, según el método de Boletín 170, p. 1–85.
K-Ar: Acta Geológica Lilloana, v. 10 no. 7, p. 141–160. Comínguez, A. H., and V. A. Ramos, 1991, La estructura
Carle, R. J., O. E. Di Persia, and H. J. Belotti, 1991, Geología del profunda entre la Precordillera y Sierras Pampeanas
sector noroeste de la provincia de Formosa: Boletín de (Argentina): evidencias de la sísmica de reflexión
Informaciones Petroleras, Tercera Epoca, no. 26, p. 2–17. profunda: Revista Geológica de Chile, v. 18, no. 1, p. 3–14.
Cazau, L. B., J. Oliver Gascón, and N. Cellini, 1976, El Cortelezzi, C. R., N. A. O. de Cuttica, and J. Solís, 1979,
Subgrupo Santa Bárbara (Grupo Salta) en la porción Estudio petrográfico y edad de las rocas basálticas
338 Comínguez and Ramos

intruídas en el Subgrupo Pirgua, Grupo Salta, Provincia de Padula, E., 1959, Valorización de las discordancias en las
Salta, República Argentina: Sexto Congreso Geológico Sierras Subandinas: Boletín Técnico de Yacimientos
Argentino, Actas II, p. 31–39. Petrolífeors Bolivianos, v. 2, no. 5, p. 7–28.
Ferreiro, O., 1989, Perspectivas de la exploración petrolera en Peddy, C., and Ch. Keen, 1987, Deep seismic reflection
el Paraguay: LXXI RANE Exploración, estado actual y profiling: how far have we come?: Geophysics: The
posibles tendencias de su tecnología, Calgary, Alberta, Leading Edge of Exploration, v. 6, no. 6, p. 22–24, 49.
p. 1–17. Ramos, V. A., 1986, El diastrofismo oclóyico: un ejemplo de
Fraga, H., and A. Introcaso, 1990, Un modelo gravimétrico tectónica de colisión durante el Eopaleozoico en el noroeste
litosférico para la Subcuenca Lomas de Olmedo (Cuenca Argentino: Revista del Instituto de Ciencias Geológicas,
del Norte y Noroeste) en la provincia de Salta, Argentina: Jujuy, v. 6, p. 13–28.
Geofísica Internacional, v. 29, no. 2, p. 89–99. Ramos, V. A., and G. I. Vujovich, 1994, Terrane history of
Galliski, M., and J. Viramonte, 1985, Cretaceous paleorift in Sierras Pampeanas (Argentina): tectonic implications
northwestern Argentina: petrological approach: Comuni- (abs.): Symposium on pre-Pangea Laurentia-Gondwana
caciones, v. 35, p. 89–92. connections, Nova Scotia, p. 35.
Galliski, M. A., and J. G. Viramonte, 1988, The Cretaceous Reyes, F. C., 1959, Posición estratigráfica de las Areniscas
paleorift in northwestern Argentina: a petrologic Superiores: Yacimientos Petrolíferos Fiscales Bolivianos,
approach: Journal of South American Earth Sciences, v. 1, Boletín Técnico, v. 1, no. 4, p. 7–36.
no. 4, p. 329–342. Reyes, F. C., 1978, Algunas consideraciones sobre la posible
Gómez Omil, R. J., A. Boll, and R. M. Hernández, 1989, edad geológica del cambio de pendiente regional en las
Cuenca cretácico-terciaria del Noroeste Argentino (Grupo Sierras Subandinas del Noroeste Argentino y Sudeste de
Salta), in G. Chebli and L. Spalletti, eds., Cuencas Sedimen- Bolivia: Revista del Instituto de Ciencias Geológicas, v. 3,
tarias de la Argentina: Universidad Nacional de Tucumán, p. 7–25.
Serie Correlación Geológica, v. 6, p. 43–64. Reyes, F. C., and J. Salfity, 1973, Consideraciones sobre la
Grier, M. E., and R. W. Allmendinger, 1991, Andean reactiva- estratigrafía del Cretácico (Subgrupo Pirgua) en el
tion of the Cretaceous Salta rift, northwestern Argentina: Noroeste Argentino. V˚ Congreso Geológico Argentino,
Journal of South American Earth Sciences, v. 4, no. 4, Actas III, p. 355–385.
p. 351–372. Reyes, F. C., J. A. Salfity, J. G. Viramonte, and W. Gutiérrez,
Grier, M. E., J. A. Salfity, R. A. Allmendinger, and S. M. 1976, Consideraciones sobre el vulcanismo del Subgrupo
Montes, 1987, La estructura precuaternaria de la quebrada Pirgua (Cretácico) en el Norte Argentino: VI˚ Congreso
de la Yesera, Salta República Argentina: las relaciones con Geológico Argentino, Actas I, p. 205–223.
la paleogeografía y la orientación de la subducción: Russo, R., R. Ferello, and G. Chebli, 1979, Llanura Chaco
Décimo Congreso Geológico Argentino, Actas I, Pampeana, in J. C. M. Turner, ed., Geología Regional
p. 193–196. Argentina: Academia Nacional de Ciencias, p. 139–180
Hall, J., 1989, Base of the crust: seismological expression, Salfity, J., 1979, Paleogeología de la Cuenca del Grupo Salta
geological evolution, and basin controls, in A. J. Tankard (Cretácico–Eogénico) del Norte de Argentina: VII˚
and H. R. Balkwill, eds., Extensional tectonics and stratig- Congreso Geológico Argentino, Actas I, p. 505–515.
raphy of the North Atlantic margins: AAPG Memoir 46, Salfity, J., 1980, Estratigrafía de la Formación Lecho
p. 41–52. (Cretácico) en la cuenca andina del Norte Argentino: Salta,
Kusznir, N. J., G. D. Karner, and S. Egan, 1987, Geometric, Universidad Nacional de Salta, Publicación Especial, Tesis
thermal and isostatic consequences of detachments in Doctoral, p. 1–137.
continental lithosphere extension and basin formation, in Salfity, J., 1982, Evolución paleogeográfica del Grupo Salta
C. Beaumont and A. J. Tankard, eds., Sedimentary basins (Cretácico–Eogénico), Argentina: V˚ Congreso Latinoamer-
and basin-forming mechanisms: Canadian Society of icano de Geología, Actas I, p. 11–26.
Petroleum Geologists Memoir 12, p. 185–203. Salfity, J. A., and R. A. Marquillas, 1986, Marco tectónico y
Levin, F. K., 1971, Apparent velocity from dipping interface correlaciones del Grupo Salta (Cretácico–Eoceno),
reflections: Geophysics, v. 36, no. 3, p. 510–516. República Argentina: En Cretácico de America Latina,
Méndez, V., J. C. M. Turner, A. Navarini, R. Amengual, and Primer Simposio, p. 174–188.
V. Viera, 1979, Geología de la región noroeste, provincias Salfity, J., R. Marquillas, M. Gardeweg, C. Ramírez, and J.
de Salta y Jujuy: Dirección General Fabricaciones Militares, Davidson, 1985, Correlaciones en el Cretácico Superior del
p. 1–118. Norte de Argentina y Chile: IV˚ Congreso Geológico
Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979, Chileno, Actas IV, no. 1, p. 654–667.
Sierras Subandinas, in J. C. M. Turner, ed., Geología Taner, M. T., and R. E. Sheriff, 1977, Application of amplitude,
Regional Argentina: Academia Nacional de Ciencias I, frequency, and other attributes to stratigraphic and hydro-
p. 95–138. carbon determination, in C. E. Payton, ed. , Seismic stratig-
Moreno, J., 1970, Estratigrafía y paleogeografía del Cretácico raphy: applications to hydrocarbon exploration: AAPG
superior de la cuenca del Noroeste Argentino, con especial Memoir 26, p. 301–327.
mención a los Subgrupos Balbuena y Santa Bárbara: Turner, J. C. M., and V. Méndez, 1975, Geología del sector
Asociación Geológica Argentina, Revista v. XXV, no. 1, oriental de los Departamentos de Santa Victoria e Iruya,
p. 9–44. provincia de Salta, República Argentina: Boletín Academia
Okaya, D. A., and C. M. Jarchow, 1989, Extraction of deep Nacional de Ciencias, v. 51, no. 1-2, p. 11–24.
crustal reflections from shallow Vibroseis data using Uliana, M. A., K. T. Biddle, and J. Cerdan, 1989, Mesozoic
extended correlation: Geophysics, v. 54, no. 5, p. 552–562. extension and the formation of Argentine sedimentary
Omarini, R. H., J. A. Salfity, E. Linares, J. G. Viramonte, and S. basins, in A. J. Tankard and H. R. Balkwill, eds., Exten-
A. Gorustovich, 1987, Petrología, geoquímica y edad de un sional tectonics and stratigraphy of the North Atlantic
filón capa lamproítico en el Subgrupo Pirgua margins: AAPG Memoir 46, p. 599–614.
(Alemanía–Salta): Revista del Instituto de Ciencias Geológ- Valencio, D. A., A. Giudicci, J. A. Mendía, and G. J. Oliver,
icas, Jujuy, v. 7, p. 89–99. 1976, Paleomagnetismo y edades K/Ar del Subgrupo
Geometry and Seismic Expression, Cretaceous Salta Rift System, Northwestern Argentina 339

Pirgua, provincia de Salta, República Argentina: VI˚ Authors’ Mailing Addresses


Congreso Geológico Argentino, Actas I, p. 519–525.
Vilela, C. R., 1967, El petróleo en las cuencas de Orán y Metán A. H. Comínguez
(provincia de Salta): II˚ Jornadas geológicas Argentinas, Gerencia de Estudios Especiales - Piso 13
Actas III, p. 425–438. Yacimientos Petroliferos Fiscales
Wernicke, B., and P. G. Tilke, 1989. Extensional tectonic Av. Diagonal Norte 777
framework of the U.S. central Atlantic margin, in A. J. 1364 Buenos Aires
Tankard and H. R. Balkwill, eds., Extensional tectonics and Argentina
stratigraphy of the North Atlantic margins, AAPG Memoir
46, p. 7–22. Victor A. Ramos
Departamento de Geología
Universidad de Buenos Aires
Ciudad Universitaria–Pabellór 2
1428 Buenos Aires
Argentina

Appendix
FIELD WORK AND PROCESSING CHARACTERISTICS

Vibroseis Data lution prior to the velocity analysis; (2) use of the
Claerbout (1976) finite difference migration algorithm; and
The section of the line where three Mertz vibrators (3) complementing the fault analysis with color complex
were used (CDPs 4000–5900 in Figure 2) involved 48 demodulation techniques (Taner and Sheriff, 1977).
recording channels per shot point with equal 100-m
receiver group intervals and a 100-m displacement from Explosive Source Data
shot to shot along 100 km of the line (24 fold). We were
able to combine two separate field programs into a single Sources involved 13–16 kg dynamite charges were
geometry simulating a unique regional experiment used in the sector defined from CDP 3076 to 4000 in the
because their respective field parameters coincided. A northern sector of seismic line 04196 (see Figure 2). Field
14–56 Hz 16-sec linear upsweep was used, while a field work consisted of 96 recording channels per shot point
record of 21 sec was collected at a sample rate of 4 msec. with equal 50-m receiver group intervals and a 50-m
The “self-truncating” extended correlation algorithm displacement per shot. Thus, fold was 48. The record was
(Okaya and Jarchow, 1989; Cominguez and Ramos, 1991) 6 sec long with a 2-msec sample rate (although a 4-msec
was used to cross correlate between the sweep and the resample rate was later used for economy).
record. The original frequency band of 14–56 Hz was Prior to channel stacking, a deconvolution operation
preserved for the first 4 sec of the seismic section, with with unmodified phase (zero phase deconvolution) was
the upper frequency decreasing from 4 sec at a rate of applied, which resulted in flattening of the amplitude
2.625 Hz/sec. The frequency band corresponding to spectrum along the 16–50 Hz frequency band. With this
different time levels of the Vibroseis seismic section is operation, precise seismic velocity analysis was ensured
summarized in Table A-1. up to 6 sec of trace length (in turn, a good acoustic
A 16-sec section was originally calculated. However, response from the crust was found to occur in this
despite careful monitoring work intended to remove sector). Likewise, since phase characteristics of the
noisy zones in each correlated trace of the field gathers, diffracted signals were kept unmodified, they could be
we were unable to image acoustic horizons in the deeper properly focalized by the migration process.
parts of the crust because of the relative signal decay on Special techniques such as finite difference migration
the traces. Consequently, it was decided to reduce the and complex demodulation were used in a similar way
final migration to 9 sec. to that of the section sector where vibrators were used.
Other important aspects of the Vibroseis data A summary of field parameters is presented in Table
processing were (1) application of “zero phase” deconvo- A-2 for both Vibroseis and explosive data.
340 Comínguez and Ramos

Table A-1—Frequency Bandwidth at Various Time Levelsa Table A-2—Lomas de Olmedo Field Parameters

Time Bandwidth Line 04196, CDP = 4000–5900


(sec) (Hz) Source: . . . . . . . . . . . . . . . . . . . .Vibroseis
0–4 14–56 Type: . . . . . . . . . . . . . . . . . . . . . .Mertz-9
5 14–53 Number: . . . . . . . . . . . . . . . . . . .3
6 14–51 Sweeps: . . . . . . . . . . . . . . . . . . . .16–24/vibrator
7 14–48 Sweep frequency . . . . . . . . . . . . .14–56 Hz, linear
8 14–45 Vibrator time: . . . . . . . . . . . . . . . .16 sec
9 14–43 Recording time: . . . . . . . . . . . . . .21 sec
Recording system: . . . . . . . . . . . .MDS-10
aSelf-truncating extended correlation was applied to the recorded Vibroseis data (line
Number of channels: . . . . . . . . . .48
04196). No consideration was given to frequency absorption, which is likely to take
place in any crustal medium.
Sampling interval: . . . . . . . . . . . .4 msec
Station Spacing: . . . . . . . . . . . . . .100 m
Near offset: . . . . . . . . . . . . . . . . .300 m
Far offset: . . . . . . . . . . . . . . . . . . .2600 m
Fold: . . . . . . . . . . . . . . . . . . . . . . .24
Spread configuration: Symmetric (2600x300-P-
300x2600)

Line 04196, CDP = 3076–4000


Source: . . . . . . . . . . . . . . . . . . . .Dynamite
Charge: . . . . . . . . . . . . . . . . . . . .13/16 kg
Source depth: . . . . . . . . . . . . . . .16/31 m
Charge interval: . . . . . . . . . . . . . .50 m
Recording time: . . . . . . . . . . . . . .6 sec
Recording system: . . . . . . . . . . . .MDS-10
Number of channels: . . . . . . . . . .96
Sampling interval: . . . . . . . . . . . .2 msec
Station Spacing: . . . . . . . . . . . . . .50 m
Near offset: . . . . . . . . . . . . . . . . .300 m
Far offset: . . . . . . . . . . . . . . . . . . .2650 m
Fold: . . . . . . . . . . . . . . . . . . . . . . .48
Spread configuration: Symmetric (2650x300-P-
300x2650)
Cretaceous Rifting, Alluvial Fan Sedimentation,
and Neogene Inversion,
Southern Sierras Pampeanas, Argentina

C. J. Schmidt C. H. Costa
Department of Geology C. E. Gardini
Western Michigan University
Kalamazoo, Michigan, U.S.A. Departamento de Geología
Universidad Nacional de San Luis
San Luis, Argentina
R. A. Astini
Facultad de Ciencias Exactas
Físicas y Naturales P. E. Kraemer
Universidad Nacional de Córdoba Facultad de Ciencias Exactas Físicas y Naturales
Córdoba, Argentina Universidad Nacional de Córdoba
Córdoba, Argentina

Abstract

T wo north-trending, west-verging, fault-bounded Neogene basement uplift systems (Sierras Chicas of


Córdoba and Serranías Occidentales of San Luis) of the Sierras Pampeanas of central Argentina are
inverted Early Cretaceous rifts. Their geometry and position 2000 km from the Atlantic continental margin and
the geometry of Neogene inversion is dependent on the earlier fabric of the basement rocks. The trends of reac-
tivated faults in the rifts are consistent with an Early Cretaceous extension direction orthogonal to the Atlantic
spreading center. The principal north-northwest rift trends were produced by dextral-oblique rifting along
previous basement sutures, and isolated depocenters may have formed as transtensional pull-apart basins. The
Sierras Chicas are the easternmost of the Pampean uplifts. They were uplifted along the eastward-dipping
Punilla thrust fault zone. Three Cretaceous depocenters containing two depositional megasequences and
volcanics are preserved as remnants of a larger basin. The sediments were deposited in restricted half-grabens
dominated by alluvial fans and playa lakes. Paleocurrent analyses indicate that the Punilla fault was a normal
fault during deposition.
Neogene inversion of normal fault trends thrusted proximal fanglomerates over their former source terrain.
Cretaceous rocks on the hanging wall of the Punilla fault zone were folded into a west-verging monocline in
the Sierra de Pajarillo area. The steep limb of the monocline is underlain by a fault-bounded wedge of cataclas-
tically deformed basement rocks.
The Serranías Occidentales of San Luis are similar to the Sierras Chicas of Córdoba. Depositional environ-
ments are similar, and fault-bounded depocenters can be identified within the larger Cretaceous San Luis
basin. The Cretaceous normal faults follow basement fabric. Neogene inversion of the Serranías Occidentales
produced short-cut faults and back-thrusts, a vertical thrust-bounded Cretaceous section (at Sierra Quijadas),
and a dramatic change of trend (north-northwest to northeast) in the basement thrust faults (Sierra del
Gigante).

Resumen

D os sistemas serranos (Sierras Chicas de Córdoba y Serranías Occidentales de San Luis), pertencientes a
las Sierras Pampeanas Orientales de Argentina y formados por basamento ascendido a lo largo de
fallas con rumbo meridional y vergencia oeste durante el Neógeno, son parte de un rift Cretácico inferior
invertido. Su geometría y ubicación a 2000 km del margen Atlántico, junto a la geometría de la inversión
neógena, dependen de la fábrica previa existente en las rocas del basamento. Los rumbos de las fallas de rift
invertidas son consistentes con la extensión ocurrida en el Cretácico inferior con direcciones ortogonales con
respecto al centro de divergencia centro-atlántico. El rumbo nor-noroeste fue producido por rifting dextral
oblicuo a lo largo de antiguas suturas del basamento, que condicionaron la generación de depocentres aislados
como cuencas transtensionales de tipo “pull-apart.” Las Sierras Chicas constituyen los contrafuertes más orien-

Schmidt, C. J., R. A. Astini, C. H. Costa, C. E. Gardini, and P. E. Kraemer, 1995, Cretaceous 341
rifting, alluvial fan sedimentation, and Neogene inversion, southern Sierras Pampeanas,
Argentina, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 341–358.
342 Schmidt et al.

tales de las Sierras Pampeanas y fueron elevadas por el corrimiento Punilla, buzante al este. Remanentes de tres
depocentros cretácicos se preservan en las Sierras Chicas, en donde se registran dos megasecuencias con
vulcanitas intercaladas. Las sedimentitas fueron depositadas en hemigrávenes restringidos, donde predomi-
naron abanicos aluviales y barreales salinos. Las paleocorrientes indican que el lineamiento Punilla se comportó
como una falla normal durante la sedimentación.
La inversión neógena del fallamiento normal cretácico sobrecorrió fanglomerados proximales sobre su
antigua área de aporte. Sucesiones cretácicas en el bloque superior del corrimiento Punilla fueron plegadas
conformando un monoclinal con vergencia oeste en el área de la Sierra de Pajarillo. El flanco empinado del
monoclinal se apoya sobre una cuña fallada de basamento cataclásticamente deformado.
Las Serranías Occidentales de San Luis son similares a las Sierras Chicas de Córdoba, desarrollando durante
el Cretácico paleoambientes análogos. En San Luis peuden también reconocerse depocentros menores contro-
lados por fracturas. Las fallas normales cretácicas están controladas por la fábrica del basamento. La inversion
Neógena de las Serranías Occidentales produjo retrocorrimientos y corrimientos de atajo en el bloque inferior
(“short-cut faults”), una sección verticalizada limitada por fallas en la Sierra de Las Quijadas y un marcado
cambio de rumbo de nor-noroeste a noreste en la orientación de los corrimientos de basamento en la Sierra del
Gigante.

INTRODUCTION
The effects of Early Cretaceous crustal extension
during the initial breakup of Gondwana were felt far
inland from the continental margin in Argentina. The
general trend of these rift basins were oblique to the
Atlantic spreading center (Figure 1) (Gordillo and
Lencinas, 1967a,b, 1979; Yrigoyen, 1975; Criado Roqué et
al. 1981; Uliana et al., 1989). However, the details of
rifting are obscure because of limited surface exposure
and the lack of subsurface information.
The purpose of this paper is to describe the pattern of
Neocomian rifting and Neogene inversion along the
Sierras Chicas of the Córdoba Province and the Serranías
Occidentales of the San Luis Province and to examine the
extent to which basement fabric has influenced later
deformation. We also consider why the rifts penetrated
the South American plate so far inland and why it has an
orientation oblique to the trend of the Atlantic spreading
center. The study is principally field based. Cretaceous
rifting was examined mainly through the provenance of
the nonmarine rift fill and the identification of probable
Cretaceous faults. Neogene inversion was examined
through field mapping and some kinematic studies.
The inverted Cretaceous rift basins of the southern
Sierras Pampeanas are located in the Sierras Chicas and
the Serranías Occidentales of San Luis (Figure 2).
Neocomian sediments in the Serranías Occidentales of
San Luis were first described in detail by Flores (1969),
who interpreted these hills and the Neogene Beazley
basin (Figure 2) as a small (15,000 km2) Cretaceous conti-
nental depocenter, which is referred to as the San Luis
basin (Flores, 1979; Riccardi, 1988) (Figure 1). The
geometry of the eastern basin in the Sierras Chicas is
inferred from isolated outcrops scattered over a wide
area within the range.
The effect of Andean deformation on the southern
Sierras Pampeanas has been outlined by Gordillo and Figure 1—Location of known or inferred Neocomian rift
basins in Argentina and Chile. Locations of study areas are
Lencinas (1979), Criado Roqué et al. (1981), Jordan and
shown in boxes. C, Córdoba; SL, San Luis; SJ, San Juan.
Allmendinger (1986), and Introcaso et al. (1987). In Terranes: CH, Chilenia; PC, Precordillera; WP, western
general, the major Neogene thrusts are spaced 50–100 Pampean; EP, eastern Pampean; RP, Rio de la Plata
km apart. They strike approximately north-south and are craton; PEL, Pelotas. (After Uliana et al., 1989; terrane
mostly west-verging. Anomalous trends do exist, boundaries after Ramos, 1988.)
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 343

Figure 2—Location map showing the principal ranges of


the Sierras Pampeanas with locations of more detailed
maps (Figures 3, 13) shown. Shaded areas are the inferred
pre-Neogene limits of Lower Cretaceous continental
deposits. (After Jordan et al., 1989.)
Figure 3—Regional geologic map of the Sierras de
Córdoba showing the three outcrop areas of Cretaceous
rocks in the eastern part of the range (Sierras Chicas) and
however, and the Serranías Occidentales and several of the locations of detailed maps (Figures 5, 9, 10). Town
the faults of the Sierras Chicas system provide excellent abbreviations: SR, Santa Rosa; LC, LaCumbre; C, La
examples of trends that depart from a north-south Calera; Cos, Cosquín; LG, La Granga; CD, Capilla del
direction. One of the goals of this paper is to examine Monte. (From Gordillo and Lencinas, 1979.)
possible causes for these anomalous trends.
The Neogene structure is fundamentally basement
involved. Where upper Paleozoic or Mesozoic rocks are rized by Dalla Salda (1987). He divided the southern
exposed on the hanging walls of the thrusts, they are Pampean ranges into (1) a western zone (Sierra de
sharply folded over the uplifted basement blocks, similar Valle Fertíl–Serranías Occidentales of San Luis) of low-
to the basement-cored folds of the Rocky Mountain to medium-grade rocks containing relatively few
foreland of the western United States (see Jordan and granites and migmatites and (2) a wide eastern zone
Allmendinger, 1986). Where no sedimentary strata are composed principally of Hercyno-type low pressure,
exposed, the prominent middle–upper Paleozoic erosion high temperature metamorphic rocks with abundant
surface is usually tilted toward the east from 2˚ to 12˚, a granitoids. Mylonites are common in both zones.
fact that is most commonly attributed to rotational uplift Martino et al. (1993) concluded that most of the
along listric faults in the basement (Jordan and mylonitic belts are Ordovician–Devonian in age and
Allmendinger, 1986; Costa, 1992). Where Paleozoic– are superimposed on upper Proterozoic high-grade
Tertiary rocks are exposed on the footwalls of these rocks. The mylonites appear to be uniformly eastward
thrusts, they are rotated to steep dips against the thrusts dipping with a top-to-the-west sense of shear. Major
and are locally overturned. Neogene thrust faults appear to follow mylonite belts
The character of the upper Precambrian–lower or boundaries between high-grade and low-grade
Paleozoic crystalline basement rocks has been summa- rocks (Baldo et al., 1993).
344 Schmidt et al.

THE SIERRAS CHICAS


Neogene Structure
The Sierras Chicas are bounded on the west by the
eastward-dipping basement-involved Punilla thrust fault
zone, which separates the range from the adjacent
Punilla Valley (Figure 3). The fault zone consists of
numerous linear overlapping fault segments with
different strikes and dips (Massabié and Szlafsztein,
1991). In this sense, it resembles ubiquitously segmented
faults that bound the major ranges of the Laramide
Rocky Mountains of the western United States. Strikes
vary from 320˚ to 20˚ and dips from about 25˚ to at least
50˚ (Kraemer et al., 1988; Massabié and Szlafsztein, 1991).
Hanging wall rocks of the Punilla fault zone consist of
upper Precambrian and lower Paleozoic metamorphic
basement and subordinate Cretaceous sedimentary rocks
(Figure 3). In most places, the strike and dip of the fault
zone parallels that of foliation in the basement rocks
(Gordillo and Lencinas, 1969; Kraemer et al., 1988;
Massabié and Szlafsztein, 1991; Kraemer and Martino,
1993).
Uplift of the Sierras Chicas along the fault zone is at
least as old as the locally derived fanglomerate of the
Casagrande Formation (Pliocene) and has continued into
the Quaternary. Footwall rocks are generally Eocene
(Cosquín Formation) (Lencinas and Timonieri, 1968;
Kraemer and Martino, 1993) to middle Pleistocene
(Kraemer et al., 1988) in age. Where exposed in the
footwall, Pleistocene sediments tend to be undisturbed
and horizontal, whereas the Tertiary section (such as at
Cosquín) dips steeply and is itself faulted over Pleis-
tocene sediments. Kinematic analysis of a north-trending Figure 4—Generalized stratigraphic columns: (a) Sierra de
segment of the fault at Cosquín is consistent with east- Pajarillo succession, (b) Estancia del Rosario succession
west (reverse slip) shortening (Kraemer and Martino, at LaCumbre, (c) Saldán Formation, and (d) Sierra de Los
Condores succession. Distance between sections is
1993), although a similar analysis at Santa Rosa (Figure 3)
shown in kilometers. Approximate relative vertical position
indicates a history of heterogeneous movement with of sections is shown. Columns a, b, and d are truncated by
both sinestral and thrust slip events (Kraemer et al., erosion at their tops. Division between upper and lower
1988). megasequence is at the top dashed correlation line
between columns. See Figure 3 for locations.
Cretaceous Sedimentary Rocks
were part of an extensive nonmarine synrift fill that
Paleogeographic Setting extended from the southern Sierras Chicas to the
Cretaceous sedimentary rocks crop out extensively in Guasayán Hills (Figure 2). The Neocomian age of the
parts of the Sierras Chicas. The Sierra de Los Condores Cretaceous deposits in the Sierras Chicas is constrained
on the south and the Sierra de Pajarillo on the north by K-Ar dates from mafic lava flows. Seventeen dates
(Figure 3) are areas of extensive outcrops on the hanging from flows in the Sierra de Los Condores Group range
wall of the Punilla thrust fault zone. Isolated outcrops from 114 ± 5 to 129 ± 8 Ma (Gordillo and Lencinas, 1967b;
also occur on the hanging wall of the Punilla fault zone Stipanicic and Linares, 1975), and the basalt within the
south of the Sierra de Pajarillo near La Cumbre (Figure Rosario conglomerate near La Cumbre is dated as 119 ± 5
3). Another large outcrop area of Cretaceous rocks occurs Ma (Gordillo and Lencinas, 1967b). Six dates for basalt
at La Calera east of the Punilla fault zone along the dikes at the Embalse de los Molinos north of the Sierra de
eastern front of the range (Figure 3). The outcrops at La Los Condres area range from 122 ± 10 Ma to 151 ± 10 Ma
Calera (Saldán Formation) are on the hanging wall of a (Stipanicic and Linares, 1975), suggesting that rifting may
steeply east-dipping reverse fault (La Calera fault) have begun in the latest Jurassic.
(Gordillo and Lencinas, 1979). These outcrops all contain The general pattern of sedimentation was controlled
conglomerates, sandstones, argillaceous material, and by the development of fault-bounded half-grabens. A
evaporites that we attribute to deposition in alluvial fan, regional survey of the stratigraphy suggests that these
braided stream, mud flat, and playa lake settings that rocks were deposited in two megasequences, although
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 345

Figure 5—Geologic map of the


Sierra de Pajarillo and Capilla del
Monte areas of the Sierras
Chicas. Lines of cross section
(Figure 6) are shown. See Figure
3 for locations. (Modified from
Pastore and Methol, 1935.)

nowhere is the complete succession present because of areas to the west and south (Uliana et al., 1989). In the
postdepositional Andean uplift and erosion. The lower late Paleozoic–Jurassic, the Sierras Chicas and other
megasequence is dominated by locally derived basement Córdoba ranges (Figure 3) apparently maintained a
detritus, whereas the upper megasequence is interlay- positive tectonic relief compared to the areas toward the
ered with mafic volcanic and volcaniclastic rocks east and west (see Jordan et al., 1989).
(Figure 4). Stratigraphic columns for the four principal
outcrop areas indicate that the two megasequences can Sierra de Pajarillo
generally be correlated from place to place (Figure 4). The northernmost outcrops of Cretaceous rocks in the
They also suggest that the three areas containing lower Sierras Chicas cover an area of about 400 km2, including
megasequence rocks (Figure 4, sections a, c, d) were the Pajarillo, Copacabana, and Maza ranges north of
separate depocenters within the larger rift basin, as Capilla del Monte (Figures 5, 6). These outcrops plunge
compared to areas such as La Cumbre (Figure 4, section below the Neogene Salinas Grande basin (Figure 3) on
b) which have a thin lower megasequence. the north, so the original extent of the basin is unknown.
These synrift continental deposits rest noncon- These sedimentary rocks are assigned to the lower
formably on upper Precambrian and lower Paleozoic megasequence (Figure 4). They were deposited in
basement rocks. It appears that intraplate rifting in the alluvial fan and playa lake environments in a small half-
Sierras Chicas was largely restricted to the Neocomian graben under arid climatic conditions (Pezzi and Astini,
and that it had no Triassic or Jurassic precursor as in 1992; Astini et al., 1993) (Figure 7). Deposits range from
346 Schmidt et al.

Figure 6—Structural cross sections from the Sierra de Pajarillo and Capilla del Monte areas. See Figure 5 for locations.

coarse, proximal, immature conglomerates to distal,


laterally extensive siltstones and mudstones with inter-
layered evaporites. The mapped alluvial fan depositional
systems on the western and southern boundaries of the
basin correlate with finer grained bajada and sporadic
braidplain deposits in the eastern and northern tracts.
Together with the distribution of evaporites on the east
and northeast, they define the asymmetry of the basin.
Paleocurrent analysis (Figure 8) indicates provenance
areas to the southwest and south across fault controlled
northwest-southeast and east-west oriented basin
margins. Conglomeratic layers were initially widespread
but gradually retreated toward the western and southern
borders, while sandy ephemeral stream and playa lake
deposits characterized the eastern tracts.
South of the Sierra de Pajarillo along the western front
of the Sierras Chicas, isolated Cretaceous outcrops have Figure 7—Block diagram of the general depositional style
survived Neogene uplift. The largest of these, east of La and structure of the Sierra del Pajarillo–Capilla del Monte
Cumbre (Figure 5), shows an interlayered volcaniclastic area. Facies: (1) upper fan debris flows, (2) middle and
upper megasequence similar to that of the Sierra de Los distal fan, (3) braided and ephemeral streams, and (4)
Condores (see Gordillo and Lencinas, 1967b). The lower muddy and saline playa lakes.
megasequence at La Cumbre is only 30 m thick
compared to at least 650 m in the Sierra de Pajarillo basin
(Figure 4, sections a, b). The following discussion is derived largely from Poiré
et al. (1988a,b) and Sánchez et al. (1990, 1993). The
Sierra de Los Condores Embalse Río Tercero Formation makes up the lower
Gordillo and Lencinas (1967a, 1979) described the megasequence and contains immature boulder to pebble
synrift volcanic rocks in this area and provided five grade conglomerates overlain by trough cross-bedded
stratigraphic sections that define the stratigraphy. The sandstones, siltstones, and bioturbated mudstones that
sequence is over 250 m thick and is referred to as the Los contain mudcracks and thin layers of gypsum. Pebble
Condores Group, which is composed of the Embalse Río imbrication and cross-bed azimuths are northeast-
Tercero Formation, the Cerro Colorado volcanics, the directed, indicating a basement terrane provenance from
Cerro Libertad conglomerates, and the Rumipalla lava the southwest across the downthrown side of the
flows (Figure 4). northwest-trending Punilla fault zone (Figure 9). The
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 347

Figure 9—General geologic map of the Sierra de Los


Cóndores area. See Figure 3 for location. (Modified from
Gordillo and Lencinas, 1967a; Sánchez et al., 1990; Sisto et
al., 1993.)

volcanic and volcaniclastic material (Figure 4). Sánchez et


al. (1990) indicate a dominantly eastern source during
second megasequence deposition, away from a west-
Figure 8—Paleocurrents plotted as rose diagrams (see
dipping normal fault boundary. They also suggest that
Nemec, 1988). Current data at locations 1, 2, 3, 9, 10, 11,
and 12 are from cobble imbrication; location 4 from imbri- the northern boundary of the basin was a down-to-the-
cation, trough cross strata, and cut-and-fill structures; south normal fault during deposition.
location 5 from planar cross strata and troughs in cross-
bedded conglomerates; location 6 from trough cross Saldán
strata, ripple lamination, and minor imbrication; location 7 The easternmost outcrops of Neocomian continental
from ripple lamination and climbing ripples; and location 8 deposits span about 150 km along the eastern flanks of
from ripple lamination, climbing ripples, and ripple trains.
the Sierras Chicas from Rio Tercero northward
(Data for the Sierra de Pajarillo from Astini et al., 1993.)
(Figure 3). They are bounded on the west by the La
Calera fault zone (Figures 3, 10). The complete succession
alluvial fans appear to have prograded toward the east along the Río Suquía (Figure 10) is over 350 m thick with
and northeast. Scarcity of braided fluvial deposits and a covered interval between the two megasequences
sand flat facies suggest an inland arid depositional envi- (Figure 4). The lower megasequence has an aggrada-
ronment with extensive bajadas. tional, slightly coarsening-upward trend with a
The Embalse Río Tercero Formation is capped with a pronounced lateral interfingering of facies. The upper
30-cm-thick horizon of purple tuffites. The overlying megasequence is less evident and partially covered, but
Cerro Libertad conglomerate of the upper megasequence shows a better organization of deposits. Piovano and
is polymictic with minor sandy trough cross-bedded Astini (1990) identified four distinct facies in the lower
conglomerates locally containing large amounts of megasequence which they compared with arid climate
348 Schmidt et al.

Figure 10). However, the amount of tectonic influence on


sedimentation is less in the upper sequence, where
paleocurrents indicate a more widespread dispersion
(locations 4, 6, Figure 10).

Neogene Inversion
Evidence suggests that the Cretaceous basins of the
Sierras Chicas were structurally inverted along the
previous synsedimentary normal faults. Our discussion
will focus on the Capilla del Monte area (Sierra de
Pajarillo and area to the south), with a briefer discussion
of the evidence for inversion of the other basins.

Geologic Setting of the Sierra de


Pajarillo-Capilla del Monte Area
The structural geology of the Capilla del Monte area,
described by Pensa (1970), Kull and Methol (1979), and
Massabié (1982), is dominated by the Neogene Punilla
fault zone that borders the western front of the Sierras
Chicas (Figure 5). South of the town of Capilla del
Monte, the fault zone trends north-northwest (350˚) and
has basement rocks and subordinate outcrops of Creta-
ceous rocks in the hanging wall. North of Capilla del
Monte, the fault zone changes to a northwest trend
(310–315˚), where it forms the southeastern boundary of
the Sierras de Pajarillo, Copacabana, and Maza. Creta-
ceous rocks of the Sierra de Pajarillo basin are on the
hanging wall of the thrust fault zone. The footwall
contains surficial Quaternary deposits and basement
rocks. We believe that considerable thicknesses of
Tertiary sediments may occur below the Quaternary
succession, as it does in footwall exposures in the Punilla
valley at Cosquín 45 km to the south (Figure 3).
The basement consists of foliated quartz feldspathic
gneiss locally interlayered with muscovite schist and
amphibolite. These foliated rocks are cross-cut by largely
unfoliated tonalitic bodies and granite. The gneiss is a
Figure 10—General geologic map of the area of Saldán high-grade rock thought to be the product of late
Formation outcrops in the eastern Sierras Chicas. Paleo- Precambrian–early Paleozoic Brasiliano metamorphism
current data at locations 1, 2, and 3 are from imbrication; (Dalla Salda, 1987). The tonalitic rocks have Late
locations 4, 5, and 6 from ripples and ripple trains in plan Devonian–Early Carboniferous ages (Stipanicic and
view. See Figure 3 for location. (Map modified from Santa Linares, 1975), suggesting a relationship to the Chánic
Cruz, 1972.)
orogeny.
The gneiss and associated rocks contain a prominent
alluvial fan and playa lake complexes. One characteristic north-northwest trending schistosity that dips 30–50˚ to
of the lower megasequence in the Saldán Formation is the northeast (Figure 11). This prevailing fabric in the
the absence of a well-developed fluvial facies. The abrupt basement rocks, described by Massabié (1982), is associ-
change from coarse-grained to fine-grained facies is ated with isoclinal folds and local mylonitization and has
typical of terminal bajadas where stream-dominated been injected by numerous aplitic bodies.
facies are poorly developed. However, the upper
megasequence contains cross-bedded conglomerates and Neogene Thrust Faults at Capilla del Monte
sandy conglomerates typical of more humid braided The Neogene Punilla fault zone in the Capilla del
alluvial systems. Monte area is characterized by an abrupt change of strike
The coarsest conglomerates of the Saldán Formation, (~45˚) (Figure 5). Cretaceous rocks are found only in the
with marble boulders up to 3.5 m in diameter, occur in hanging wall. With respect to the Cretaceous rocks, the
the lower megasequence adjacent to the La Calera fault, north-northwest trending segment of the fault zone is
suggesting that it was down to the east during deposi- exposed at a slightly deeper structural level than the
tion. Paleocurrent indicators for the lower megasequence northwest-trending segment that bounds the Sierra de
show an eastward dispersal pattern (locations 1, 2, 3, 5, Pajarillo basin. The two segments are connected by a 7-
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 349

minerals along foliation planes. The fault zone follows


the regional trend of foliation, but in many places
foliation departs from being parallel to the fault trend.
Nevertheless, foliation planes of nearly all orientations in
the fault-bounded block show evidence of shear. Appar-
ently, slip on fractures of many different orientations
parallel to foliation accommodated westward rotation of
the basement block between the two thrust faults. The
basement rocks on the hanging wall of the upper, more
easterly fault are also fractured, but they are considerably
less cataclastically deformed than the basement between
the faults, and chloritic alteration along foliation planes is
virtually absent beyond a few tens of meters away from
the fault. Analysis of slip on fractures at two different
locations in the upper, more easterly fault zone suggests
a direction of shortening oriented nearly west-east,
implying mostly dip slip with subordinate sinestral slip
on the main fault (Figure 12a, b). Slip on fractures in the
basement rocks 100–500 m east of the fault zone are
compatible with this same shortening direction
(Figure 12c).
A similar pattern of faulting and monocline develop-
ment in Cretaceous rocks exists east of La Cumbre
(section D–D', Figure 6). Gordillo and Lencinas (1967b,
Figure 11—Contoured lower hemisphere stereoplot of profile 1) interpreted Cretaceous rocks to be present on
poles to foliation in basement rocks on the hanging wall of the footwall of the lower, more westerly thrust.
the Punilla thrust in the Capilla del Monte area (Figure 5). However, there is no direct evidence that any Cretaceous
Mean foliation attitude shown as a great circle.
rocks exist on the footwall of the thrust.
The northwest-trending pair of Neogene thrust faults
km-long zone of faults that cut through the Uritorco that bound the Sierra de Pajarillo have the same
granite on a north-northeast trend (Figure 5). geometry as those described above except that they are
Both major segments of the Punilla fault zone are exposed at a shallower structural level, almost entirely
similar in that they are each composed of two principal within the Cretaceous deposits (section A–A', Figure 6).
faults 1.2–1.7 km apart. The fault on the east is better As elsewhere, there is a west-dipping monoclinal panel
exposed in each area and forms a steep mountain front. It of proximal fanglomerates between the two Neogene
dips 45–60˚ east or northeast. The fault on the west is thrusts, and the hanging wall of the upper (more
poorly exposed through Quaternary cover, but it locally easterly) thrust contains flat-lying Cretaceous rocks that
cuts pediments north of Capilla del Monte. The structure become progressively finer grained away from the
of the Cretaceous deposits between the two faults is best frontal thrust. Dip separation for the upper thrust is
exposed south of Capilla del Monte in the hill, locally about 100 m, whereas dip separation for the lower one is
called the Cabeza de Soldado, east of the town of Los unknown but probably greater.
Cocos (C–C', Figure 6), and north of Capilla del Monte
Inversion and Control of Fault Trends at
along the Quebrada de Luna (A–A', Figure 6).
Capilla del Monte
At Cabeza de Soldado (section C–C'), very coarse
proximal debris flows dip toward the west at 35–45˚ In every area studied in detail along the trend of the
between the two faults (Figure 6). About 1 km upslope Punilla fault zone, the hanging wall of the more westerly
on the hanging wall of the more easterly fault, outcrops thrust contains the most proximal Cretaceous rocks,
of this conglomerate are finer grained and rest with hori- generally massive pebble to boulder conglomerates with
zontal nonconformity on the basement. The geometry of crude stratification. Analysis of cobble and boulder
these Cretaceous outcrops suggests a westward-dipping imbrication (lower megasequence) at the Cabeza
monocline with thrust faults at the upper (anticlinal) and Soldado (location 10, Figure 8) indicates a western source
lower (synclinal) hinges. Dip separation on the thrust perpendicular to the trace of the Punilla fault zone. Two
that cuts through the upper hinge is unknown, but is other locations on the hanging wall (locations 11, 12,
probably relatively small (a minimum of about 50 m). Figure 8) also show a predominance of eastward-
Dip separation on the lower more westerly thrust is also oriented paleocurrent directions. Although no statistical
unknown, but may be as great as several kilometers. study was made at the La Cumbre site, examination of
Exposures of basement within the fault-bounded imbrication at three locations indicates a western source
basement block (section C–C') show strong cataclasis for the lower megasequence conglomerates. At two
with closely spaced (2–10 cm) fault-parallel fractures in locations along the northwest-trending faults bordering
unfoliated rocks and ubiquitous foliation-parallel shear the Sierra de Pajarillo (locations 1, 9, Figure 8), paleocur-
associated with chloritic alteration of ferromagnesian rent analysis of cobble imbrications gives a direction
350 Schmidt et al.

Figure 12—Faults and striae


orientations and P and T axis
scatter plots or contoured
diagrams for slickensided
surfaces at (a and b) two
locations along the upper fault
of the Punilla thrust zone and (c)
one location 100–500 m east of
the fault. See Figure 5 for
locations.

toward the north-northeast, approximately normal to the higher than the highest basement peak (Uritorico at
strike of the fault zone. 1950 m). A minimum dip separation of 1500 m is implied.
The paleocurrent data and the proximal nature of the Speculation that this east-trending fault zone, and its
hanging wall conglomerates of the more westerly thrust parallel counterpart in the basement rocks about 5 km
of the Punilla fault zone clearly suggest that this thrust farther south, were active during Cretaceous deposition
was the bounding normal fault for the Sierra de Pajarillo is supported by analyses of paleocurrent directions for
basin and for the deposits of the Sierras Chicas to the the Cretaceous succession on the north side of the fault
south and that it was reactivated during the Neogene. (locations 2–5, Figure 8). They indicate a southern source
There was probably little or no Cretaceous deposition to perpendicular to this fault zone (Astini et al., 1993). The
the west of the Punilla fault zone. deposits against the fault zone at Los Terrones are
Like the western boundary faults of the Punilla fault composed of middle to proximal fan facies, suggesting
zone, the east-west trending fault that borders the Sierra close proximity to an uplifted basement source. This fault
de Pajarillo basin on the south is clearly reactivated. The zone may have been an intrabasinal structure during
fault zone at this contact is nearly vertical with the south rifting and acted to isolate the Pajarillo basin on the north
side up. The Cretaceous rocks at Los Terrones (Figure 5) from the rest of the half-graben (Figures 7, 8). This fault
on the north side of the zone are locally turned up zone may have been a tear fault parallel to transport
against this fault and dip northward at 20˚–30˚ (section during Neogene shortening, thus permitting the
B–B', Figure 6). Dip separation is unknown, but the base basement block on the south to move vertically more
of the Cretaceous rocks on the south side is projected than the block containing the Pajarillo basin on the north.
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 351

The evidence of approximately east-west shortening SERRANIAS OCCIDENTALES OF


in this area and at Cosquin (Kraemer and Martino, 1993)
suggests that the northwest-trending segment of the
SAN LUIS
Punilla fault zone bordering the Sierra de Pajarillo had a The Serranías Occidentales of the San Luis Province
significant component of sinistral slip during the are a suite of small mountain ranges northwest of the city
Neogene. of San Luis that form the southernmost ranges of the
According to Massabié and Szlafsztein (1991), there is Sierra de Valle Fertil system (Figures 2, 13). This section
a strong tendency for the Punilla fault zone to follow briefly summarizes the Neogene structure, the Creta-
foliation trends in the Sierras Chicas. We have observed ceous sedimentary rocks, and the evidence for Neogene
that the north-northwest trending, northeast-dipping inversion of these ranges.
foliation orientation between Capilla del Monte and La
Cumbre is parallel to the main fault zone, although local Neogene Structure
changes in foliation are cut by the main fault. About 5
km due east of Capilla del Monte and east of the The general structure of the Serranías Occidentales of
Uritorico granite outcrops, foliation within the basement San Luis is similar to that of the Sierras Chicas, except
changes abruptly from north-northwest to northwest that much more of the Cretaceous section is preserved
before the basement rocks dive beneath the Cretaceous and less basement is exposed. The Neogene structure of
cover (Figure 5). The northwest-trending foliation is each of the ranges is essentially that of a west-facing
subparallel to the northwest-trending boundary faults of monocline of Cretaceous rocks above basement. The
the Sierra de Pajarillo. This suggests that Neocomian monoclinal structure appears to be nearly continuous
extension exploited an earlier basement fabric and that over the entire 200 km length of the ranges. The west
the Neogene faults exploited the Cretaceous faults. (synclinal) hinge of the monocline is broken by an east-
dipping Neogene thrust that trends north-northwest
Other Evidence of Inversion in the Sierras Chicas following the general trend of the Valle Fertil. The trend
South of Capilla del Monte in the Saldán area changes abruptly to north-northeast in the southern
(Figure 3), the western boundary of the Cretaceous Sierra del Gigante (location 1, Figure 13). Rocks on the
Saldán Formation is everywhere a steep eastward- hanging wall of the thrust are basement (schists,
dipping reverse fault (La Calera fault) with a minimum gneisses, and mylonites) and usually conglomeratic
of 50 m of dip separation (Gordillo and Lencinas 1967a, Cretaceous rocks.
1979). At La Calera, two reverse faults are separated by a The Cretaceous rocks on the frontal monocline above
few hundred meters (Figure 10) (Santa Cruz, 1972). the thrust have dips (30˚–60˚ W) similar to comparable
Because of the existence of very proximal facies with structures in the Sierras Chicas. Locally, however, such
locally derived marble boulders up to 3.5 m in diameter as at the western border of Sierra Quijadas, the Creta-
on the hanging wall of this fault and because of the ceous conglomerates are vertical above the thrust.
eastward-oriented paleocurrents for the fan facies Footwall rocks consist of Cenozoic clastics. The Tertiary
(Figure 10), it seems apparent that the La Calera fault section below the thrust, such as along the western
zone was reactivated with a reversal of throw. Inversion boundary of Sierra del Gigante, is steeply west dipping
has produced a uniform eastward 5˚ dip for the Saldán to overturned. Average dip on the bounding thrust is
Formation. 50–55˚ E, and movement is mostly dip slip. However, at
The nature of inversion in the Sierra de Los Condores the best exposure of this fault, striking 345˚ and dipping
has yet to be fully documented, although the basin has 52˚ E in western Sierra del Gigante (location 2, Figure 13),
clearly been inverted by movement on the Punilla fault slickenlines show a small sinistral component of
zone (e.g., Sisto et al., 1993). Similar to the Sierra de movement consistent with a nearly east-west direction of
Pajarillo and northern Sierras Chicas, the Punilla fault shortening (Figure 14).
zone at the Sierra de Los Condores is composed of two The basement rocks in Sierra Guayaguas are bounded
principal faults about 1–1.5 km apart (Figure 9). on the east by a west-dipping thrust (location 5, Figure
However, only basement rocks are exposed between the 13) that we interpret as a back-thrust of the principal
two faults. Little is known about these faults. The more east-dipping thrust. The footwall rocks are conglomer-
easterly fault may be a west-dipping back-thrust of ates, basalts, and mudstones that are turned up to
Neogene age that is antithetic to the principal fault vertical below the thrust. These have been dated as Late
(following Gordillo and Lencinas, 1967a), or it may be an Triassic from faunal evidence (Bossi, 1975; Bossi and
east-dipping Cretaceous normal fault, making the Bonaparte, 1978). The conglomeratic Cretaceous section
principal fault a Neogene short-cut fault. Although lies with slight angular discordance above these rocks,
Gordillo and Lencinas (1967a) show the fault on the indicating that this area was the site of proximal deposi-
eastern boundary of the Sierra de Los Condores basin to tion in a Cretaceous basin that was superimposed on
be a west-dipping normal fault, Sisto and Cortés (1992) proximal deposition in a Triassic basin.
and Sisto et al. (1993) suggest that the most recent The west-dipping thrust along the eastern border of
movement on this fault is oblique with reverse and the ranges (Figure 13) is speculative because it is not
sinestral components. They also show that minor east- exposed. However, such a thrust is required along the
dipping thrusts locally follow the strike of the west- eastern border of Sierra del Gigante where the prominent
dipping principal fault (Figure 9). basement outcrops of the range end abruptly at a sharp
352 Schmidt et al.

northwest-trending photolineament. The thrust fault


mapped west of the ranges in the Bermejo basin (Rio
Desaguadero fault) has been traced to the subsurface
west of Sierra Guayaguas where it has been interpreted
as the major Neogene thrust responsible for uplifting
these ranges and the Sierra de Valle Fertil to the north
(Snyder et al., 1990).
Several northeast-trending faults separate the
Serranías Occidentales into their respective ranges. Most
of these faults probably have minor Neogene movement.
However, the fault that crosses the northern part of the
Sierra del Gigante (location 4, Figure 13) is associated
with minor folds in the basement rocks on its northern
side, suggesting at least two separate episodes of fault
movement, one of which may have occurred in Creta-
ceous time. Likewise, the normal fault on the southern
boundary of the basement rocks of the range (location 3,
Figure 13) shows evidence of Cretaceous movement but
no movement during the Neogene.

Cretaceous Sedimentary Rocks


The Cretaceous rocks are well exposed in each of the
ranges of the Serranías Occidentales and have been pene-
trated by at least six wells in the Neogene Beazley basin
to the southeast (Figure 2). Continental deposits are
confined to the north-northwest elongated San Luis
basin, which is composed of several depocenters and is
about 3000 km2 in area (Figure 2). The stratigraphy has
been described elsewhere (Flores, 1969, 1979; Flores and
Criado Roqué, 1972; Criado Roqué et al., 1981; Manoni,
1985; Riccardi, 1988). The Lower Cretaceous stratigraphy
(El Gigante Group) was described from the southern
Sierra del Gigante (Flores and Criado Roqué, 1972), and
the formation names have been applied to the rocks in
the Sierra Guayaguas and Sierra Quijadas (e.g., Bossi and
Bonaparte, 1978; Rivarola and Di Paola, 1992) (Figure 15).
Correlation between ranges is tenuous, and it is clear that
each range constitutes a local depocenter and has a
slightly different stratigraphy.
The El Gigante Group (Figure 15) is composed of two
sequences of conglomerates (Los Riscos and La Cruz) that
grade laterally and vertically into fluvial sandstones and
sandy conglomerates (El Jume and Toscal formations),
which in turn grade laterally and vertically into lacustrine
argillites and gypsum beds of the La Cantera Formation
(Figure 15). The cumulative thickness of the group is
estimated to be 1670 m (Yrigoyen, 1975). Although it has
been described as a continuous continental succession
(e.g., Manoni, 1985) resembling the lower megasequence
of the Sierras Chicas, no physical connection between the
two rift basins is known to have existed.
The highest levels of the La Cruz Conglomerate in the
Sierra Quijadas are interlayered with basalt flows that are
thought to be equivalent to the Cretaceous basalts in the
Sierras Chicas (Gordillo, 1972). They have yielded K-Ar
radiometric ages of 106 ± 5 Ma to 161 ± 3 Ma (Oxfordian–
Albian) (González, 1971; Yrigoyen, 1975; Llambías and
Figure 13—General geologic map of the Serranías Occi- Brogioni, 1981). The palynology of the La Cantera
dentales of San Luis. Numbers refer to locations described Formation indicates an Early Cretaceous age (Flores,
in the text. (Modified from Yrigoyen, 1981.) 1969; Yrigoyen, 1975; Bonaparte, 1981).
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 353

Figure 14—Faults and striae


orientations and P and T axis
scatterplot for four measure-
ments of slickensided
surfaces in the principal
thrust fault zone at location 2
(Figure 13), western Sierra del
Gigante.

Little detailed sedimentologic work has been done on composed of the same lithologies as the footwall
the El Gigante Group. However, sedimentologic studies basement rocks. In addition, imbrication of tabular clasts
of the different formations (Manoni, 1985; Rivarola and Di indicates an eastward paleocurrent trend; the conglomer-
Paola, 1992) indicate depositional settings and paleoenvi- ates fine away from the fault. The footwall basement
ronments similar to those of the Cretaceous deposits in block was probably the source of the clasts in the La
the Sierras Chicas: alluvial fans and bajadas with a small Cruz, suggesting that this eastern fault was the original
braidplain and playa lakes in an arid climate. Cretaceous normal fault boundary separating the
The Lagarcito Formation (up to 250 m thick) contains Creatceous conglomerate fan from its basement source.
a basal conglomerate-mudstone-gypsum association Because the Neogene thrust does not appear to have
overlain by red sandstone. These have been interpreted reactivated the normal fault, the thrust is interpreted as a
as playa lake deposits grading upward to higher energy short-cut fault.
fluvial deposits (Manoni, 1985). Although assigned a The La Cruz unconformably overlies the larger mass
Late Cretaceous age (Yrigoyen, 1975), vertebrate fossils in of basement exposed to the east. This mass is thrusted
the sandstones suggest an Early Cretaceous age eastward over Upper Triassic rocks along its eastern
(Bonaparte, 1970, 1971; Sánchez, 1973). These rocks are border (location 5, Figure 13). The basement rocks of this
more widespread than the El Gigante Group (Manoni, larger mass are different from the western fault-bounded
1985) and are attributed to a postrift sag stage of defor- sliver, consisting of quartz feldspathic gneisses, schists,
mation and deposition. and amphibolites with an extensively developed
mylonitic fabric (Simon and Rossello, 1990; Castro de
Neogene Inversion Machuca and Carrizo, 1993). The fact that no mylonitic
basement clasts occur in the La Cruz Conglomerate
Evidence that the western boundary thrust of the indicates that the source of the clasts was probably the
Serranías Occidentales was a down-to-the-east normal nonmylonitic basement to the west.
fault during the deposition of the Cretaceous rocks is It is likely that basement fabric controlled the orienta-
reasonably well documented for each of the four ranges tion of the Cretaceous normal fault and the Neogene
of the system. Furthermore, in the Sierra Guayaguas and faults (frontal thrust and back-thrust). Although foliation
Sierra del Gigante, the fabric of the basement rocks can attitudes vary considerably within the basement masses,
be shown to have influenced the trend of both Creta- the strong mylonitic fabric of the main basement mass is
ceous normal faulting and Neogene thrust faulting. parallel to the normal fault on the west (345˚, 70˚ E). In
addition, the western sliver of basement does not have
Sierra Guayaguas the same mylonitic fabric, suggesting that the normal
A critical observation for the inversion history of this fault follows a fundamental north-northwest trending
range is that two masses of basement rocks in the boundary in the basement (i.e., between mylonitic and
northern part of the range are separated by a narrow belt nonmylonitic rocks). Simon and Rossello (1990) reported
of Cretaceous conglomerates (probably the La Cruz that the kinematic indicators in the mylonite are consis-
Conglomerate). The western mass, composed mainly of tent with largely dip-slip motion (top to the west) with a
quartzites and amphibolites, is a fault-bounded sliver 300 minor sinestral component. The age of the movement
m wide and 2 km long (location 6, Figure 13). The associated with the mylonitization is unknown, but
western boundary fault is the principal east-dipping Simon and Rossello (1990) and Martino et al. (1993)
Neogene thrust. The eastern boundary fault separates the suggest a Paleozoic age.
basement block from the La Cruz Conglomerate. The The steep eastward dips of the mylonitic zones on the
conglomerate dips 35–40˚ W into the fault, and the fault western side of the main basement mass give way to
dips steeply toward the east (70–80˚). The conglomerate moderate westward dips on the eastern side of the mass.
on the hanging wall adjacent to the fault is made up of In fact, the attitude of the mylonitic fabric on the eastern
proximal debris flows with clasts up to 2 m in diameter side on the hanging wall of the back-thrust (location 5,
354 Schmidt et al.

Sierra del Gigante


Sierra del Gigante is composed of a basement massif
flanked on the north and south by proximal facies of
minor Cretaceous basins. The western boundary is a
Neogene thrust that places the basement mass and the
flanking basins over Tertiary sediments. The principal
structure of the basement rocks is a large east to east-
northeast striking antiform (Figure 13) of low to medium
grade metasedimentary rocks (Gardini and Costa, 1987;
Gardini, 1993). Foliations and compositional layering dip
40–50˚ SE in the southern limit and 40–50˚ N in the north.
The northern Cretaceous outcrops (La Cruz Conglom-
erate) generally dip gently northward away from the
basement massif, and cobble imbrications indicate
transport from the southwest. We suggest that the
northeast-trending, northwest-dipping fault (location 4,
Figure 13) was down on the northwest during the depo-
sition of the La Cruz Conglomerate in this area. Minor
folds in the foliated rocks of the fault zone suggest that
the most recent movement (probably Neogene) was
right-reverse. The western border of the Cretaceous
outcrops is a west-facing monocline bounded by the
principal east-dipping thrust. These outcrops, like those
in Sierra Guayaguas and Sierra Quijadas, have cobble
Figure 15—Generalized stratigraphy of the Cretaceous imbrications that suggest transport from the west. It
rocks in the Serranías Occidentales of San Luis. Numbers appears that this northern basin was bounded by normal
in parentheses are maximum thicknesses in meters. The faults on the west and on the south, both of which
Quebrada del Barrio Formation is only in Sierra Guayagus. underwent Neogene movement.
Lagarcito Formation may be Lower Cretaceous. See text The basement massif in the main part of the range is
for discussion. (Modified from Yrigoyen, 1975.) thrust westward over Tertiary rocks. Geologic maps of
the range (e.g., Gardini, 1993) suggest that the fabric of
the metamorphic rocks is cut at nearly right angles by the
Figure 13) is nearly identical to that of the back-thrust, thrust. However, we find that along the western border
suggesting that the back-thrust was also localized along of the range, the foliation in the basement rocks turns
the earlier fabric. abruptly southward in less than 100 m from the thrust
and becomes parallel with it in several places. The fault-
Sierra Quijadas
parallel foliation contains numerous parasitic (z) folds
Although no basement rocks are exposed in Sierra that plunge steeply west-northwest. The fabric is not
Quijadas, the morphology of the Neogene uplift is clear, cataclastic as might be expected if its orientation were
and the facies changes in the Cretaceous rocks provide changed by folding associated with movement along the
reasonable indirect evidence of inversion. The Creta- Neogene thrust. Rather, cataclasis in the basement rocks
ceous outcrops are part of a west-verging anticline that is is confined to within a few meters of the thrust. The
nearly flat-topped between an eastern and western abrupt change in foliation attitude appears to be related
hinge. The western hinge marks the abrupt transition to an earlier ductile folding event of probable Paleozoic
from horizontal strata to vertical strata above the origin; the thrust, and presumably the Cretaceous
boundary thrust. The eastern hinge is a more gradual normal fault as well, followed this earlier fabric.
change from horizontal to gently eastward-dipping Where the western boundary thrust changes trend
strata. abruptly to the southwest in the southern part of the
Inversion is inferred from very rapid facies changes in range (location 1, Figure 13), the hanging wall rocks are
the La Cruz Conglomerate from the western border to the Cretaceous La Cruz Conglomerate. They are folded
the eastern anticlinal hinge (a distance of about 4 km). In into a northwest-vergent anticline above the thrust. The
this distance, the La Cruz can be seen to coarsen upward structure and stratigraphy is similar to that at Sierra
in the section and fine to the east, from proximal debris Quijadas, with coarse conglomerates near the thrust
flows with clasts over 3 m in diameter to braided stream fining eastward and down section into sandstones,
deposits with only a few pebbly zones. Clast imbrication mudstones, and gypsum beds. Cobble imbrication is
in the conglomerates suggests a western source. It seems scattered but indicates transport mainly from the west or
likely that the western boundary thrust was a normal northwest. We believe that the thrust was once a down-
fault during deposition of the La Cruz and that the to-the-southeast normal fault during deposition of the
uplifted basement block on the footwall was the Cretaceous rocks. The conglomerate facies appear to be
sediment source for the deposit. middle to distal fan deposits, suggesting a westward
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 355

recession of the Cretaceous mountain front and the and a bimodal provenance. In both areas, the lower
possibility that more proximal facies are now on the megasequence has a short fining-upward succession
footwall of the Neogene thrust. covered by a thick upward-coarsening succession that
The southern boundary of the basement massif is a we interpret as deposition during a long period of quies-
northeast-trending normal fault that dips 50˚ SE (location cence after a brief period of active, fault-controlled subsi-
3, Figure 13). The conglomerates are proximal clast- dence (see Blair and Bilodeau, 1988). An upper succes-
supported debris flows with the largest clast size sion (Lagarcito Formation) representing a sag phase of
exceeding 2 m. The conglomerates fine rapidly south- deposition is present in the San Luis basin, which may
eastward, and cobble imbrications in the middle and correlate with the rocks of the upper megasequence from
proximal fan facies indicate transport from the north. the Sierras Chicas.
These rocks were assigned to the lower conglomeratic In the Sierras Chicas rift system, comparatively little of
sequence (Los Riscos) by Flores (1969) (Figure 15). Near the original geometry and none of the presumed inter-
the fault contact, they are horizontal or dip gently into connectivity or yoking of depocenters is preserved. The
the fault. Away from the fault, they dip 10–15˚ SE, Pajarillo depocenter has a bimodal sediment transport
defining a broad hanging wall roll-over undisturbed by and at least one important cross fault. The Los Condores
Neogene thrusting. Foliation in the gneisses of the depocenter shows a reversal of sediment transport
footwall dips parallel to the normal fault, strongly direction from the west-bounding fault to the east-
suggesting basement fabric control. This normal fault is bounding fault between the lower and upper megase-
on strike with the northeast-trending segment of the quences. Nevertheless, the geometry of the three
Neogene thrust that bounds the range (location 1, Figure principal preserved depocenters seems relatively simple
13). Apparently the abrupt change of trend of the compared to documented examples of tensional rifts
boundary thrust at this location (from north-south to (e.g., Leeder and Gawthorp, 1987) and transtensional
northeast-southwest) was influenced by the normal fault, systems (e.g., Nilsen and McLaughlin, 1985). Perhaps
and the portion of the normal fault southwest of location this is because the level of exposure permits only the
1 was reactivated as a thrust. Where the normal fault earliest fault-sediment patterns to be examined.
changes to a thrust, proximal Riscos conglomerates on However, the simplicity may be real because of the
the hanging wall of the normal fault grade into the more fundamental control exerted by the west-verging Punilla
distal overlying La Cruz Conglomerate on the hanging and La Calera fault zones that follow earlier basement
wall of the thrust. fabric. Consequently, a westward asymmetry may have
The frontal Neogene thrust that bounds the Serranías dominated the fault-sediment pattern all along the
Occidentales continues southward from Sierra del Sierras Chicas.
Gigante into the Cerrillada de La Cabra (Figure 13). The surface exposures of the Serranías Occidentales
Rocks interpreted to be Cretaceous in age (probably represent less than 25% of the known surface area of the
Lagarcito Formation) are found in scattered outcrops on San Luis basin. In spite of documented cross faults and
the footwall of the thrust. Although the Lower Creta- bimodal sediment transport directions, our surface
ceous rocks appear to be confined to the region east of observations are biased toward the overwhelming
the Cretaceous normal fault that became the range importance of the western boundary fault system that
bounding thrust, the Upper(?) Cretaceous fluvial section follows the basement fabric all along the ranges. Future
(sag facies) appears to continue to the west of the fault examination of seismic data for the other 75% of the
here. According to Manoni (1985), seismic and well data basin will probably reveal a more complex fault-
indicate that the Lagarcito continues westward to the Rio sediment pattern. Nevertheless, the importance of the
Desaguadero fault, which is its sedimentary depositional reactivated west-verging western boundary fault system
or erosional limit. is established, and we suggest that a dominant westward
asymmetry will eventually be revealed for this basin in
seismic sections.
In the San Luis basin, Cretaceous rifting was superim-
DISCUSSION AND CONCLUSIONS posed on earlier Triassic rift basins in the north and
south (Criado Roqué et al., 1981; Uliana et al., 1989), but
Although no physical connection has been demon- in the Sierras Chicas, there is no evidence for rifting
strated, the two north-south elongate Cretaceous rift before the Neocomian. This observation may be
basins in the southeastern Sierras Pampeanas have important to petroleum potential and future exploration
several characteristics in common. Besides nearly south in the Levalle basin and Macachin trough
identical ages, the similarities include (1) a westward (Figure 1). These subsurface basins are directly in line
asymmetry with east-dipping normal faults dominating with the Sierras Chicas rift trend. If these subsurface rifts
the rift basin geometry and depositional setting; (2) a are the uninverted continuation of the basins of the
lower depositional sequence dominated by proximal Sierras Chicas, they may not contain sediments older
debris flows grading upward and laterally into mid-fan, than Neocomian. Furthermore, because the sediments
braidplain, and playa lake deposits; (3) lava flows at the were deposited under arid conditions without significant
top of the lower megasequence or base of the second lacustrine facies, the source rock potential of these basins
megasequence of deposition; and (4) the presence of may be limited. It is possible, however, that these unin-
depocenters characterized by intersecting fault trends verted basins do have a younger lacustrine facies that has
356 Schmidt et al.

Neogene contractional features follows many, but not all,


of the normal fault trends. Several faults, including the
northeast-trending normal fault at Sierra del Gigante,
were not favorably oriented for reactivation by east-west
contraction. South of about the latitude of San Luis
(~33˚ S lat), the Sierras Chicas rift trend is in the subsur-
face (Levalle basin) and is not inverted, and the
Serranías–San Luis rift trend has only minor inverted
ranges. This latitude corresponds with the southern
boundary of the flat subduction corridor of the Nazca
plate (Jordan et al., 1983) confirming that contractional
effects of convergence affected Pampean basement weak-
nesses principally within that corridor.
As Jordan and Allmendinger (1986) have described,
the Neogene fault and uplift geometry is similar to that
Figure 16—Principal trends of Cretaceous normal faults at
Capilla del Monte and Sierra del Gigante, showing inferred of the Late Cretaceous–Paleogene Rocky Mountain
slip directions and direction of extension compatible with foreland of the western United States. The similarities
the fault orientations. can be extended to include the folding of the cover rocks
as well, in which asymmetric folding of cover accompa-
nies cataclastic deformation in an underlying fault-
been stripped off the Sierras Chicas because of inversion. bounded wedge of basement.
In both of the rift basins, normal faults are controlled
by the anisotropy of the basement fabric (principally
foliation in gneiss and mylonite). At Capilla del Monte in Acknowledgments We thank the Fulbright Commission,
the Sierras Chicas and at Sierra del Gigante in the Mobil Oil Corporation, Marathon Oil Corporation, and the
Serranías Occidentales, we can identify six fault orienta- Faculty Research Fund at Western Michigan University for
tions reflected in the basin configuration (Figure 16). Of financial assistance in this cooperative work and the Universi-
these, only one (the east-west orientation at Capilla del ties of Córdoba and San Luis for logistical support. We
Monte) is not known to be parallel to the local basement acknowledge the very helpful discussions of various aspects of
fabric. The direction of crustal extension most compatible this work with María Sánchez, David Rivarola, Roberto
with all of these fault trends is west-northwest to east- Martino, Blaine Hall, Art Slingsby, Hayden Tanner, Carlos
southeast (Figure 16), orthogonal to the trend of the Early Barcat, and Steve Derksen. Francisco Quintana provided a
Cretaceous spreading center (Figure 1) reported by most photogeologic interpretation of the Capilla del Monte area that
authors (e.g., Fairhead and Binks, 1991; Tankard et al., greatly helped our interpretation. We are especially grateful for
1995). This implies that there is some strike slip the helpful advice and able field assistance of Bruce Malamud
component to most of the normal faults and that the and Gustavo Zulliger. Schmidt would like to thank Teresa
main north-northwest trending fault zones in each basin Jordan and Richard Allmendinger for their help in getting him
were subjected to right-oblique rifting. The isolated started with work in the Pampean ranges. Analysis of fault-slip
depocenters in each rift trend may have formed as data was facilitated with software (Fault-kin, version 3.25)
transtensional pull-apart basins controlled by basement written and provided by R. W, Allmendinger, R. A. Marrett
fabric (see also Fernandez-Seveso and Tankard, 1995). and T. Cladouhos. We also used Stereonet, version 4.5, by
Although the details of the internal geometry of the Allmendinger to plot foliation data. Tony Tankard, Craig
rifts appear to be controlled by the local basement fabric, Knutson, Teresa Jordan, and Art Slingsby provided helpful
it is unlikely that this fabric alone was responsible for the editorial comments and reviews. Some of the figures were
extensional deformation affecting the crust so far inland. prepared by Bob Havira, Bill Lozier, and Matt Malin. Bev
The regional trend of the basement fabric, however, must Britt, Lisa Choiniere, and Robbie Zenero typed most of the
reflect fundamental zones of crustal weakness. In the manuscript.
Sierras Chicas, this is likely the Late Proterozoic (Brasil-
iano orogenic cycle) suture between the eastern Pampean
terrane and the Rio de La Palta craton (see Ramos, 1988) REFERENCES CITED
(Figure 1). In the Serranías Occidentales, it is the Valle
Fértil lineament that is probably an early Paleozoic Astini, R. A., L. I. Pezzi, and G. A. Massei, 1993, Paleo-
(Ocloyic) suture between the western and eastern geografía y paleoambientes del Cretácico de la Sierra
Pampean terranes (see Martino et al., 1993) (Figure 1). In Pajarillo–Copacabana–Maza, noroeste de Córdoba: Actas
XII Congreso Geológico Argentino, v. 1 p. 170–176.
the Late Cretaceous, rifting along these sutures was Baldo, E. G. A., C. Schmidt, S. Bertolino, and M. Martinez,
abandoned for the more favorably oriented Late Protero- 1993, Características estructurales y petrográficas de las
zoic suture between the Rio de La Plata craton and the filitas del borde occidental de La Sierra de Pocho, Córdoba
Pelotas terrane (Ramos, 1988) (Figure 1). (abs.): IX Reunión de Microtectónica Resúmenes.
Because the rift pattern followed previous crustal Blair, T. C., and W. Bilodeau, 1988, Development of tectonic
anisotropy, it is not surprising that the pattern of cyclothems in rift, pull-apart, and foreland basin: sedimen-
Cretaceous Rifting and Neogene Inversion, Southern Sierras Pampeanas, Argentina 357

tary response to episodic tectonism: Geology, v. 16, p. Gordillo, C. E., and A. Lencinas, 1967a, Geología y petrología
517–520. del extremo norte de la Sierra de Los Cóndores, Córdoba:
BonaBonaparte, J. F., 1970, Pterodaustro guinazui, pterosaurio Boletín Academia Nacional Ciencias de Córdoba, v. 46,
de la Formacion Lagarcito, Provincia de San Luis, p. 73–108.
Argentina y su significado en la geología regional: Acta Gordillo, C. E., and A. Lencinas, 1967b, El basalto nefelínico
Geológica Lilloana, v. 10, p. 209–225. de El Pungo: Boletín Academia Nacional Ciencias de
Bonaparte, J. F.,1971, Descripción del Cráneo y mandíbulas de Córdoba, v. 46, p. 110–115.
Pterodaustro guinazui (Pterodactyloidea–Pterodaustriidae Gordillo, C. E., and A. Lencinas, 1969, Pefil geológico de la
nov. ) de la Formación Lagarcito, San Luis, Argentina. Sierra Chica de Córdoba en la zona del río Los Molinos,
Museo de Ciencias Naturales Mar del Plata, Publicación 1, con especial referencia a los diques traqui-basálticos que la
p. 263–272. atraviesan: Boletín Academia Nacional Ciencias de
Bonaparte, J. F.,1981, Los Fósiles Mesozoicos, in Geología y Córdoba, v. 47, 25–50.
Recursos Minerales de la Provincia de San Luis: Relatorio Gordillo, C. E., and A. Lencinas, 1979, Sierras Pampeanas de
Octavo Congreso Geológico Argentino, p. 97–99. Córdoba y San Luis, in Geología Regional Argentina:
Bossi, G. E., 1975, Geología de la cuenca de Marayes–El Academia Nacional de Ciencias de Córdoba, v. 1,
Carrizal, Provincia de San Juan, Republica Argentina: p. 577–650.
Actas VI Congreso Geológico Argentino, v. 1, p. 23–38. Introcaso, A., A. Lion, and V. Ramos, 1987, La estructura
Bossi, G. E., and J. F. Bonaparte, 1978, Sobre la presencia de un profunda de las Sierras de Córdoba: Revista Asociacion
diosaurio prosaurópodo en La Formación Quebrada del Geológica Argentina, v. 42, p. 177–187.
Barro, en el borde austral de la cuenca de Marayes–El Jordan, T. E., and R. W. Allmendinger, 1986, The Sierras
Carrizal, Triásico superior de San Juan: Acta Geológia Pampeanas of Argentina: a modern analogue of Rocky
Lilloana, v. 15, p. 41–47. Mountain foreland deformation: American Journal of
Castro de Machuca, B., and M. Carrizo, 1993, Contribución al Science, v. 286, p. 737–764.
conocimento del cerro Guayaguas, Provincia de San Luis Jordan, T. E., B. L. Isacks, R. W. Allmendinger, J. A. Brewer, V.
(abs.): IX Reunión de Microtectónica Resúmenes. A. Ramos, and C. J. Ando, 1983, Andean tectonics related
Costa, C., 1992, Neotectónica del sur de la sierra de San Luis: to geometry of subducted Nazca plate: GSA Bulletin, v. 94,
Ph.D. dissertation, Universidad Nacional de San Luis, San p. 341–361.
Luis, Argentina, 390 p. Jordan, T. E., P. Zeitler, V. Ramos, and A. Gleadow, 1989,
Criado Roqué, P., C. A. Mombru, and V. A. Ramos, 1981, Thermochronometric data on the development of the
Estructura e interpretación tectónica, in Recursos naturales basement peneplain in the Sierras Pampeanas, Argentina:
de la Provincia de San Luis: Relatorio Octavo Congreso Journal of South American Earth Sciences, v. 2, p. 207–222.
Geológico Argentino, p. 155–192. Kraemer, P. E., and R. D. Martino, 1993, La falla de la Sierra
Dalla Salda, L., 1987, Basement tectonics of the southern Chica, Cabalgamiento de basamento sobre una cuña
Pampean Ranges, Argentina: Tectonics, v. 6, p. 249–260. sedimetaria imbricada, Cosquín Córdoba (abs.): IX
Fairhead, J. D., and R. M. Binks, 1991, Differential opening of Reunión de Mocrotectónica Resumenes.
the central and south Atlantic Oceans and the opening of Kraemer, P. E., and R. D. Martino, M. Giambastiani, Y. J.
the West African rift system: Tectonophysics, v. 187, Sfragulla, 1988, Análisis dinámico y cinemático preliminar
p. 191–203. de la falla de Santa Rosa, Departamento Calamuchita,
Fernandez-Seveso, F., and Tankard, A. J., 1995, Tectonics and Provincia de Córdoba: Actas V Reunión de Microtectónica,
stratigraphy of the late Paleozoic Paganzo basin, western v. 1, p. 107–114, Córdoba.
Argentina, and its regional implications, in A. J. Tankard, Kull, V., and E. Methol, 1979, Descripción geológica de la hoja
R. Suarez, and H. J. Welsink, Petroleum basins of South 21i (Alta Gracia): Dirección Nacional de Geología y
America: AAPG Memoir 62, this volume. Minería Boletín, Buenos Aires, no. 55, 72 p.
Flores, M. A., 1969, El Bolsón de Las Salinas en la Provincia de Leeder, M. R., and R. L. Gawthorp, 1987, Sedimentary models
San Luis: Cuartas Jornadas Geológicas Argentinas, Buenos for extensional tilt-block/half-graben basins: Geological
Aires, v. 1, p. 311–327. Society of London Special Publication 36, p. 139–152.
Flores, M. A., 1979, Cuenca de San Luis, in Segundo Simposio Lencinas, A., and A. Timonieri, 1968, Algunas características
de Geología Regional Argentina: Academia Nacional de estructurales del Valle de Punilla (Córdoba): Actas
Ciencias, Córdoba, v. 1, p. 645–769. Terceras Jornadas Geológicas Argentinas, Buenos Aires,
Flores, M. A., and P. Criado Roqué, 1972, Cuenca de San Luis, v. 1, p. 195– 208.
in A. F. Leanza, ed., Geología Regional Argentina: Llambías, E. J., and N. Brogioni, 1981, Magmatismo
Academia Nacional Ciencias, Córdoba, p. 567–579. Mesozoico y Cenozoico, in Geología y Recursos Minerales
Gardini, C., 1993, La estructura íntima del basamento de la de la Provincia de San Luis: Relatorio del Octavo Congreso
Sierra de el Gigante, San Luis: XII Congreso Geológico Geológico Argentino, p. 101–115.
Argentino, v. 3, p. 1–10. Manoni, R., 1985, Geología del subsuelo de la Cuenca del
Gardini, C. E., and C. H. Costa, 1987, Geología del basamento Beazley: Boletín de Informaciones Petroleras YPF, p. 34–46.
de la Sierra de el Gigante, Provincia de San Luis, Martino, R. D., C. Simpson, and R. D. Law, 1993,
Argentina: Decimo Congreso Geológico Argentino, Actas Taconic–(Ocloyic) aged west-directed ductile thrusts in
III, p. 43–46. basement rocks of the Sierras Pampeanas, Argentina (abs.):
González, R., 1971, Edades radimétricas de algunos cuerpos GSA Abstracts with Programs, v. 25, p. A233.
eruptivos de Argentina: Revista de la Asociación Massabié, A. C., 1982, Geología de los alrededores de Capilla
Geológica Argentina, v. 26, p. 411–412. del Monte y San Marcos, Provincia de Córdoba: Revista
Gordillo, C. E., 1972, Petrografía y composición quimica de los Asociación Geológica Argentina, v. 37, p. 153–173.
basaltos de la Sierra de Las Quidadas-San Luis-y sus rela- Massabié, A. C., and C. F. Szlafsztein, 1991, Condiciones
ciones con los basaltos cretácicos de Córdoba: Notas geomecánicas y edad del fallamiento neotectónico en Las
breves, Boletín de la Asociacion Geológica de Córdoba, v. Sierras Pampeanas orientales, Córdoba, Argentina:
1, p. 127–129.
358 Schmidt et al.

Congreso Argentino de Geología e Ingeniería Aplicada, Sisto, F., and J. Cortés, 1992. Tectónica Cretácica–Cenozoica
v. 6, p. 154–168. del tramo sur de la Sierra de Los Cóndores, Sierras
Nilsen, T. H., and R. J. McLaughlin, 1985, Comparison of Pampeanas de Córdoba: Actas de la VII Reunión de Micro-
tectonic framework and depositional patterns of the tectónica I, p. 63– 69.
Hornelen strike-slip basin of Norway and the Ridge and Sisto, F., M. Sánchez, and C. Perez Posio, 1993, Dinámica y
Little Sulphur Creek strike-slip basins of California, in K. T. cinemática Cenozoica del frente de fracturación Puesto
Biddle and N. Christie-Blick, eds., Strike-slip deformation, Viejo–La Cantera, Sierra de Los Cóndores, Provincia de
basin formation, and sedimentation: SEPM Special Publi- Córdoba, Republica Argentina (abs.): IX Reunión Microtec-
cation 37, p. 79–103. tónica Resúmens.
Nemec, W., 1988, The shape of the rose: Sedimentary Snyder, D. B., V. A. Ramos, and R. W. Allmendinger, 1990,
Geology, v. 59, p. 149–152. Thick-skinned deformation observed on deep seismic
Pastore, F., and R. Methol, 1953, Descripción geológica de la reflection profiles in western Argentina: Tectonics, v. 9, p.
Hoja 19i (Capilla del Monte, Córdoba): Dirección Nacional 773–788.
Minería, Buenos Aires, Boletín 79, 70 p. Stipanicic, P. N., and E. Linares, 1975, Catálogo de edades
Pensa, M. V., 1970, Perfil estratigráfico entre Ongamira y radimétricas determinadas para la Republica Argentina I,
Copacabana (Provinia de Córdoba): Boletín Asociación Años 1960–1974. Asociación Geológica Argentina, Publi-
Geológica Córdoba, v. 1, p. 52. cación Especial B3, p. 1–42.
Pezzi, L I., and R. A. Astini, 1992, Facies de evaporitas conti- Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M.
nentales en el Cretácico de Córdoba: Cuarta Reunión Turic, A. B. Franca, E. J. Milani, B. B. de Brito Neves, N.
Argentina Sedimentología, v. 3, p. 133–120. Eyles, H. de Santa Ana et al., 1995, Tectonic controls of
Poiré, D. G., M. L. Sánchez, and M. B. Villegas, 1988a, Facies basin evolution in southwestern Gondwana, in A. J.
sedimentarias de la sección inferior del Grupo Sierra de Tankard, R. Suarez, and H. J. Welsink, Petroleum basins of
Los Cóndores, Embalse de Río Tercero, Provincia Córdoba, South America: AAPG Memoir 62, this volume.
Republica, Argentina: Contribuciones sobre el Cretácico de Uliana, M. A., K. T. Biddle, and J. Cerdan, 1989, Mesozoic
América Latina Actas, p. 121–132. extension and the formation of Argentine sedimentary
Poiré, D. G., C. Larriestra, M. Villegas, and M. Sánchez, 1988b, basins, in A. J. Tankard, and H. R. Balkwill, eds., Exten-
Análisis morfométrico de las psefitas basales del Grupo sional tectonics and stratigraphy of the North Atlantic
Sierra de Los Cóndores (Cretácico), en el área del Embalse margins: AAPG Memoir 46, p. 599–614.
Río Tercero, Córdoba: Actas Segunda Reunión Argentina Yrigoyen, M. R., 1975, La edad Creácica del Grupo del
de Sedimentología, p. 217–221. Gigante (San Luis) y su relacón con cuencas circunvecinas:
Piovano, E. L., and R. A. Astini, 1990, Facies de abanico Actas Primero Congreso Argentino de Paleontología y
aluvial semiárido en la Formación Saldán, Quebrada del Bioestratigrafía, Tucumán, v. 2, p. 29–56.
Río Suquia, Sierra Chica de Córdoba: Actas Tercer Reunión Yrigoyen, M. R., 1981, Mapa geológico de la Provincia de San
Argentina de Sedimentología, p. 217–222. Luis y regiones circunvecinas: Relatorio Octavo Congreso
Ramos, V. A., 1988, Late Proterozoic–early Paleozoic of South Geológico Argentino, p. 7.
America—a collisional history: Episodes, v. 11, p. 168–174.
Riccardi, A. C., 1988, The Cretaceous system of southern
South America: GSA Memoir 168, 161 p.
Rivarola, D., and E. Di Paola, 1992, Paleoambientes desérticos
Authors’ Mailing Addresses
en el Potrero de La Aguada, Sierra de Las Quijadas, Christopher J. Schmidt
Cretácico de la Provincia de San Luis, Argentina: Cuarta Department of Geology
Reunión Argentina de Sedimentología, p. 79–84. Rood Hall, Room 1187
Sánchez, M. L., M. B. Villegas, and D. G. Poiré, 1990, Paleo- Western Michigan University
geografia del Cretácico inferior en el área de La Sierra de Kalamazoo, Michigan, 49008-5150
Los Cóndores, Provincia de Córdoba, Republica
Argentina: Tercera Reunión Argentina Sedimentolgía,
U.S.A.
Actas, p. 235–246.
Sánchez, M. L., C. Perez Posio, and F. Sisto, 1993, Modelo R. A. Astini
cinemático de la cuenca de los Condores (Cretácico P. E. Kraemer
inferior), in la Provincia de Córdoba, Republica Argentina Facultad de Ciencias Exactas, Físicas y Naturales
(abs.): V Journadas Pampeanas de Ciencias Naturales, Universidad Nacional de Córdoba
p. 43. Avda. Vélez Sársfield 299
Sánchez, T. M., 1973, Redescripción del cráneo y mandíbulas 5000 Córdoba
de Pterodaustro guinazui Bonaparte (Pterodactyloidea, Argentina
Pterodustriidae): Ameghiniana, v. 10, p. 313–325.
Santa Cruz, J. N., 1972, Geología al Este de la Sierra Chica
(Córdoba): Valle del Río Primero: Boletín Asociación
C. H. Costa
Geológica Córdoba, v. 1, p. 102–109. C. E. Gardini
Simon, O. W., and E. A. Rossello, 1990, Observaciones Departamento de Geología
petrológicas y cinemáticas en las metamorfitas del Cerro Universidad Nacional de San Luis
Guayaguas, San Juan, Argentina: Actas X Congreso 5700 San Luis
Geológico Argentino, v. 1, p. 464–468. Argentina
Structural Inversion and Oil Occurrence in the
Cuyo Basin of Argentina

Daniel Dellapé Andrés Hegedus


Astra C.A.P.S.A. Enterra Oil Field Rental
Buenos Aires, Argentina Neuquen, Argentina

Abstract

M ost of the hydrocarbon reserves in the Cuyo basin are contained in 15 oil fields that are mainly struc-
turally controlled. They include several pools in the 40 million m3 range. Tectonic analysis based on
seismic data tied to well control suggests that most closures relate to folds and reverse faults that are geneti-
cally associated with an earlier extensional fault system that developed during the early Mesozoic collapse of a
late Paleozoic orogenic belt. Reconstruction of the source rock (Middle Triassic Cacheuta Formation) and
reservoir paleogeography (Triassic–Tertiary Potrerillos, Río Blanco, Barrancas, and Papagayos formations)
indicates that synsedimentary extension and differential subsidence were key factors that induced an irregular
distribution of organic-rich strata and porosity development.
Cenozoic contraction linked to Andean orogenesis inverted the Triassic half-grabens and created structural
closures. This resulted in local reservoir enhancement and access to effective charge after late Cenozoic
regional migration. Prospective closures consist of elongate, irregularly spaced to en echelon anticlines and
plunging noses. Axial surfaces display eastward or westward vergence, and shallow folds are replaced at
depth by faulted structures (e.g., Tupungato, Barrancas, La Ventana, and Río Tunuyán fields).
The roots of the structural highs involve stratigraphic depocenters and high-angle faults that show normal
separation at depth and reverse separation at intermediate levels (e.g., Vizcacheras field). The amount of
inversion decreases from west to east, and in the most deformed areas, the cores of the folds were penetrated
by faulting and popped up structures shaped as bivergent thrust wedges bounded by master faults and
converging back-thrusts.

Resumen

L a mayor parte de las reservas de hidrocarburos de la Cuenca Cuyana se encuentran alojadas en 15


campos petrolíferos controlados por trampas de tipo estructural. Estos yacimientos incluyen varias
acumulaciones en el orden de los 40 millones m3. Estudios tectónicos basado en el análisis de secciones
sísmicas calibradas mediante perfiles de pozo, sugieren que la mayor parte de los cierres se relaciona con
pliegues y fallas extensionales, que se desarrolló durante el colapso eo-Mesozoico de un cinturón orogénico del
Paleozoico Tardío. Reconstrucciones de la paleogeografía correspondiente a los intervalos que contienen a la
roca madre (Formación Cacheuta, meso-Triásica) y a los reservorios (Formaciones Potrerillos, Río Blanco,
Barrancas, y Papagayos del Tiásico-Terciario), señalan que los fenómenos de subsidencia diferencial y
extensión sinsedimentaria fueron instrumentales para producir una distribución irregular de los estratos ricos
en materia orgoánica y de los desarrollos porosos.
Procesos de compresión neocenozoica ligados a la orogénesis de los Andes provocaron la inversión de los
hemigrábenes mesozoicos, dando lugar a cierres estructurales y a vías de migración que proveyeron una
efectiva carga de hidrocarburos cuando el petróleo migró en el Cenozoico más tardío. Los cierres prospectivos
se presentan como anticlinales alargados y narices buzantes, con un espaciamiento irregular y arreglo en-
echelón. Las superficies axiales muestran vergencia oriental y occidental, y las estructuras someras de tipo
pliegue son reemplazadas en profundidad por configuraciones falladas (ej. Yacimientos Tupungato, Barrancas,
La Ventana y Río Tunuyán). Las raíces de los altos estructurales involucran zonas de espesamiento de la serie
triásica, y fallas de alto ángulo que muestran separación normal en profundidad y separación inversa a niveles
intermedios (ej. Yacimiento Vizcacheras). La magnitud de la inversión decrece de oeste a este, y en la mayor
parte de las áreas deformadas, los núcleos de los pliegues se han visto penetradas por las fallas y por estruc-
turas de tipo “pop-up”, que tienen forma de cuña limitada por las fallas maestras y por zonas de retrocabal-
gamiento.

Dellapé, D., and A. Hegedus, 1995, Structural inversion and oil occurrence in the Cuyo 359
basin of Argentina, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of
South America: AAPG Memoir 62, p. 359–367.
360 Dellapé and Hegedus

Figure 1—Structural framework


of the Cuyo basin, showing the
main productive structures and
regional faults. Inset (a) shows
the configuration in latest
Cretaceous–early Tertiary time;
inset (b) shows a reconstruc-
tion at the end of Cenozoic
time.

INTRODUCTION high-angle faults. Most recent work, supported by better


seismic coverage (e.g., Turic et al., 1981; Pombo, 1986),
attributes the structures to medium- or low-angle faults
The Triassic–Tertiary Cuyo basin, which covers an involving basement. According to this model, the
area of about 30,000 km 2 in the northern Mendoza dominant faulting is reverse and eastward verging and
Province, underlies a lowland segment of the Argentine probably associated with antithetic thrusts and conjugate
foreland. It is located between the mountains of the pre- transverse faults involving strike-slip offsets. A more
Cordillera and the elevated crystalline blocks of the recent hypothesis (Legarreta et al., 1993) emphasizes
Pampean Ranges. Exploration activity during the past 50 changes in structural polarity and the presence of relay
years has delineated 15 major oil fields in the basin. transfers along the master faults. On this basis, some of
These fields are predominantly structurally controlled. the oil-bearing structures are attributed to inversion of
They are mostly half-anticlines, doubly plunging anti- Triassic extensional faults.
clines, and noses arranged in an en echelon pattern After examining the influence of structural style on
defining three main trends (Figure 1) (Criado Roqué et hydrocarbon occurrence, we review some examples in
al., 1960; Turic et al., 1981; Pombo, 1986). which the relationships between stratigraphy and
Although faults associated with the closures are structure are clear. Our aim is to develop a model that
believed to play a significant trapping role, there is no will support more effective exploration and production
general agreement on the deep configuration of the activity in the future.
dominant structural style. The traditional interpretation Our study area stretches from south of the city of
(e.g., Padula, 1972) is that the productive structures are Mendoza to the southern part of the Cacheuta subbasin,
large basement blocks bounded by eastward-verging encompassing the area of the Cuyo basin oil fields.
Structural Inversion and Oil Occurrence in the Cuyo Basin of Argentina 361

Figure 2—Schematic cross section based on seismic lines and wells. Tectonic inversion of the western trend of productive
anticlines and half-grabens is shown. See Figure 1 for location.

STRATIGRAPHIC FRAMEWORK laid down at a stage when differential subsidence was


important and allowed development of a widespread
Underlying the oil-bearing Mesozoic deposits are depocenter. Thick sedimentary columns accumulated
Paleozoic sedimentary rocks that have been overprinted locally. This depositional pattern had particular signifi-
by medium- to low-grade metamorphism, referred to as cance for the distribution of organic-rich lacustrine facies
the Villavicencio Formation (Harrington, 1941), and the in the Cacheuta Formation and for the distribution of
volcanic and pyroclastic outpourings of the Choiyoi reservoir quality sandstones of the Potrerillos, Río Blanco,
Group (Rolleri and Criado Roque, 1968). The latter unit is and Barrancas formations (Figure 2).
closely related to basal members of the Triassic succes- During the sag stage of subsidence, volcanic activity
sion (Kokogian and Mansilla, 1989). was sporadic. During the Jurassic and Early Cretaceous,
The Triassic rocks were deposited unconformably on the Punta de las Bardas Formation, a series of olivine
the Paleozoic substratum. They are genetically related to basalt flows, was extruded (Rolleri and Criado Roque,
an episode of rifting that induced block rotation and fault- 1968). By Miocene–Pliocene time, depositional rates had
controlled subsidence. During the early depositional increased considerably and a 1500- to 2200-m-thick
phase, the sediments were restricted to a series of nonmarine succession was deposited. This sequence
partially isolated depressions known as the “Superse- includes fine-grained deposits, sandstones, conglomeratic
cuencia Inferior” (Kokogian and Mansilla, 1989). In sandstones, conglomerates, and tuff interbeds known as
proximal positions, the rocks are represented by agglom- the Mariño, La Pilona, Tobas Grises, and Mogotes forma-
erates and conglomerates (Río Mendoza Formation) that tions. Subsidence patterns during the late Cenozoic are
were deposited in alluvial fan systems. In more distal attributed to tectonic loading linked to encroachment and
locations and in younger deposits, the coarse alluvial fan uplift of the Precordilleran fold and thrust belt and a
facies are replaced by finer sediments laid down in fluvial foreland dissected by the Pampean Ranges.
and lacustrine depositional systems (Las Cabras
Formation). Regional deposition was initially controlled
by fault-driven subsidence linked to rifting, followed by a
period of regional subsidence attributed to thermal decay. STRUCTURAL STYLES AND HISTORY
The upper part of the local Mesozoic succession is OF DEFORMATION
known as the “Supersecuencia Superior” (Kokogian and
Mansilla, 1989) and comprises the Potrerillos, Cacheuta, The study area is dominated by an irregularly spaced
Río Blanco, and Barrancas formations. These units series of well-defined elongate anticlines. They are locally
represent the rift to sag transition and the regional postrift broken by reverse faults that have axial planes
sag period itself. They consist of fluvial, deltaic, and lacus- displaying eastward or westward vergence. Legarreta et
trine deposits (Kokogian and Mansilla, 1989) that were al. (1993) have suggested that the principal structural
362 Dellapé and Hegedus

Figure 3—Schematic cross section showing initiation of tectonic inversion. The sedimentary wedges are not well developed.
See Figure 1 for location.

features began to develop at the same time as initiation Because the Tertiary succession is piled up without
of Triassic extension. The present-day structures were marked angular unconformities and because Pliocene
outlined after Cenozoic Andean tectonism. strata of the Mogotes Formation are involved in the
Anticlinal closures are aligned along three north- shallow folding, the period of contractional deformation
northwest striking structural trends (Figure 1). The is attributed to a fairly young age (Padula, 1972; Jordan
western series of folds includes culminations at and Ortíz, 1987).
Cacheuta, La Pilona, Estructura Intermedia, Tupungato, Seismic sections across the northern part of the Cuyo
Piedras Coloradas, Chañares Herrados, and Alto Verde. basin show definite evidence of asymmetric extensional
The central trend includes the following structures: Cruz structures during sedimentation of the Middle and
de Piedra, Lunlunta, Barrancas, Carrizal, Río Tunuyán, Upper Triassic sequences. A series of master faults
La Ventana, Vacas Muertas, Punta de las Bardas, and El induced sedimentary thickening into half-graben type
Quemado. The eastern trend is more subdued both at the tectonic depressions (Figures 3, 4, 5). The deformation
surface and in the subsurface and comprises the sequence involves three stages: (1) extension, block
Rivadavia, Zampal, and Vizcacheras structures. rotation, and half-graben fill; (2) regional sag and differ-
The individual culminations are connected through a ential subsidence; and (3) compression and inversion.
series of structural saddles. All of them show asymmetric
cross sections, with steeper flanks commonly linked to Extension, Block Rotation, and
conspicuous reverse faults. Seismic control demonstrates Half-Graben Fill
that shallow folded configurations at Tertiary and upper
Mesozoic levels are replaced at depth by less obviously During this episode from about 245 to 232 Ma,
folded strata displaying a series of differentially interval A–B was deposited (Figures 3, 4, 5). Geodynamic
displaced and rotated fault block geometries (Figures 3, control is attributed to extensional collapse of a late
4). These contrasting architectures are associated with Paleozoic orogen developed after intracratonic compres-
stratigraphic intervals that have laterally varying thick- sional deformation and crustal thickening near the
nesses. They are ultimately linked to a deformational margin of the Gondwana supercontinent (Mpodozis and
history that superimposed contractional strain on an Kay, 1990; Legarreta et al., 1993). The extensional episode
older extensionally induced basement fabric. is recorded in the suite of listric and planar faults.
Structural Inversion and Oil Occurrence in the Cuyo Basin of Argentina 363

Figure 4—Schematic stratigraphic section showing distribution of sedimentary units and sequence boundaries. Not to scale.
(Courtesy of L. Legarreta.)

Fault throws are locally as much as 3500–4000 m. Compression and Inversion


They resulted in substantial Middle Triassic depocenters
at Tupungato and Piedras Coloradas and also near During the latest Cenozoic, compressional stresses
Vizcacheras, Zampal, and Rivadavia (Figure 1). The related to Andean orogenesis induced tectonic inversion
distribution of these depocenters is coincident with the and differential uplift of some of the principal Triassic
eastern and western structural trends. The central trend depocenters, thus forming structural culminations.
developed around a paleotopographic high that Depositional troughs adjacent to the principal basin-
coincided with a transfer zone linking the opposed half- forming faults became structural highs due to east-west
graben systems. Location above the paleohigh is contraction and lateral slip. Anticlinal culminations were
reflected in attenuated development of the entire Triassic thus forced on top of former depocenters (see Williams et
succession. al., 1989).
The seismic control shows that some positive struc-
Regional Sag and Differential Subsidence tural features such as Tupungato, Piedras Coloradas, and
Vizcacheras are developed above an unusually thick
Interval C–D (Figures 3, 4, 5), deposited from about Triassic succession. As a result of the inversion process
232 to 25 Ma, records gradual reduction of rifting and (see Bally, 1984; Biddle and Rudolph, 1988; Williams et
progressive increase of regional subsidence. The Jurassic, al., 1989), originally low blocks have turned into positive
Cretaceous, and early Tertiary sequences (Barrancas, features. Typically the roots of the structural highs
Punta de las Bardas, and Divisadero Largo formations) involve an expanded stratigraphic succession, and the
appear to be only locally displaced by extensional former high-angle master faults show reverse separation
faulting and display a more regular thickness trend. at intermediate stratigraphic levels and fade out into the
However, regional isopach mapping and seismic control younger stratigraphic units.
(Figures 3, 4) indicate that the hanging wall sides of the Several examples of inverted half-graben and graben
half-graben troughs suffered a comparatively higher structures are illustrated in Figures 2 and 3. Several of
subsidence rate during the latest Triassic and these structures show opposing polarities. Those located
Jurassic–Paleogene periods. A 3000-m-thick succession at to the west along the western productive trend involve
Tupungato, for example, correlates with a 500-m-thick faults dipping to the east beneath thick Triassic deposits.
section at Río Tunuyán and with a 1000-m-thick succes- Those along the eastern structural trend behave as a
sion at Vizcacheras (Figure 5). mirror image with a sedimentary prism to the east
The effect of differential subsidence is also demon- adjacent to the master faults (Figure 3).
strated by the facies patterns within each genetic The area along the central structural trend shows
interval. Local concentrations of sandstones deposited small inverted half-grabens verging in an eastward or
by successive fluvial systems are usually coincident with westward direction, depending on their original location
originally low paleopositions, and shales with high on the former regional high. The area is characterized by
paleopositions. a thin, lower Mesozoic stratigraphic sequence that is
364 Dellapé and Hegedus

(a)

(b)

Figure 5—(a) Interpreted seismic section and (b) geologic cross section in an west-east orientation. Note the tectonic
inversion in the western part. Numbers on contacts are ages (in Ma). Symbols as in Figure 4.
Structural Inversion and Oil Occurrence in the Cuyo Basin of Argentina 365

interpreted as a transfer zone between half-graben within the 2000-m-thick Tertiary succession demon-
systems rotated in opposite directions. strates that normal faulting was not active during the
The magnitude of inversion changes laterally and Cenozoic.
shows variable expression at successive stratigraphic Well data and geologic analysis of the seismic sections
levels. Therefore, it is possible for the inversion to be support the interpretation shown in Figure 3, which
inconspicuous at basement level while significant defor- shows a general east–west thinning and locally thick
mation of strata is present at the base of the Tertiary. This Triassic sections due to normal fault offsets. Near the
implies that the post-Triassic and Tertiary successions structural culminations, the Triassic succession thins and
were deformed due to folding while the basement is devoid of reservoir quality sandstones.
reacted rigidly and lacked noticeable arching. In those
zones where inversion was milder, the sag stage Creta- Inverted Normal Faults
ceous deposits were locally shortened and flexed
without faulting (Figure 3). At more advanced stages of Inception of the compressional stress field associated
deformation, the cores of the folds were penetrated by with Andean orogenesis is reflected in reverse reactiva-
faulting and developed “pop-up” wedges flanked by tion of the inherited normal master faults (Figure 3).
downward-convergent reverse faults that are believed to Seismic images of intermediate and young stratigraphic
reflect inverted antithetic and second-order synthetic horizons demonstrate reverse separations. The newly
accommodation faulting. developed structural fabric includes back-thrusting along
Some of the master faults that were reactivated as antithetic faults, block arching, and general uplift of the
reverse faults are linked to zones of antithetic deforma- area that hosts the Vizcacheras oil pool.
tion interpreted as converging back-thrusts. These faults Due to its location distant from the mountain front,
play an important role in some of the oil-bearing struc- the Vizcacheras block suffered only relatively moderate
tures, which locally has a direct effect on hydrocarbon compressional stress compared to the patterns recorded
trapping. The presence and degree of development of for the western trend. The local structure at Vizcacheras
these structures depends on the magnitude of the local shows only a modest amount of inversion. The present
compressional stresses. The most common structural structural culmination reflects close proximity to mildly
traps are related to arching at shallow levels, while a inverted antithetic faults and dips toward a structural
reduced magnitude of fold-related closure is more low on the hanging wall side of the original half-graben.
common at depth. At greater depths, blocks that The present trapping configuration is believed to be an
preserve noninverted normal offsets are largely unex- inverted rollover feature.
plored, but appear to be affected by burial depth and
deteriorated reservoir quality.
EFFECT OF FAULTING ON THE
HABITAT OF OIL
STRUCTURAL STYLES IN THE
VIZCACHERAS BLOCK Syndepositional faulting was not only an important
factor controlling the structural evolution and final
In spite of its condensed sedimentary development, geometric configuration of many hydrocarbon traps, it
the stratigraphy of the Vizcacheras block records all of also exerted a considerable influence on source rock and
the depositional events documented in the thicker reservoir distribution in the Cuyo basin.
sections to the west. One similarity is the thick sandstone
sections deposited in depressions (Figure 3) during the Source Rock Distribution
thermal sag stage (Barrancas and Papagayos formations).
An analysis of seismic sections and well control demon- The persistence of lacustrine systems throughout
strates significant variations in fault activity. Triassic deposition favored organic productivity and
preservation. The main source rock interval of the Cuyo
Normal Faults basin is in the Cacheuta Formation. Thinner and less
organic-rich members also occur in the Las Cabras,
Normal faults formed boundaries for early Mesozoic Potrerillos, and Río Blanco formations. The area near La
depocenters. The oldest Triassic sequences were Pilona, Tupungato, Piedras Coloradas, and Chañares
deposited on the downthrown side of the faults, while Herrados represents the best hydrocarbon “kitchen” in
the upturned flanks of the blocks remained above depo- the basin because of its thick, organic-rich shales and
sitional base level. These normal faults have dip separa- adequate thermal history (Chebli et al., 1984; Jordan and
tions on the order of 1000 m. Ortíz, 1987). The region near Vizcacheras and Rivadavia
Differential block faulting was active during depo- is less favorable because of its reduced thickness and less
sition of the Late Triassic and younger Mesozoic than optimum level of organic maturity.
sequences. Displacement rates were subdued but still Comparison of lithofacies distributions and faulting
sufficient to induce lateral changes in facies and in thick- patterns shows that listric faulting created depositories
nesses. An essentially parallel set of seismic reflections adjacent to the fault plain that supported preservation of
366 Dellapé and Hegedus

lacustrine facies under euxinic bottom conditions. Hydrocarbon Recovery


During lacustrine highstands, clastic influx was reduced
and persistent organic sedimentation occurred on the Petroleum generation in the basin is a comparatively
downthrown sides of the half-grabens. Burial under a recent process, suggesting that some of the master faults
thick Cenozoic succession matured the lacustrine shales may have acted as conduits for hydrocarbon migration.
and charged the Mesozoic reservoirs (Chebli et al., 1984; For many years it was believed that zones producing at
Jordan and Ortíz, 1987; Rosso et al., 1987). commercial rates were naturally fractured reservoirs,
such as the Río Blanco pools in Tupungato, Piedras
Reservoir Development Coloradas, and Chañares Herrados fields. These fields
have produced oil for nearly 50 years. However, full-
Commercial oil pools in the Cuyo basin are contained diameter core studies and considerable horizontal
mostly in the sandstone reservoirs deposited by fluvial drilling have demonstrated a negligible amount of frac-
systems. Stratigraphic reconstructions suggest the turing in oil-productive zones. Besides the presence of
existence of a drainage network dominated by major sandstones, the most important factor controlling reser-
distributories running parallel to the half-graben axes. voir quality is the magnitude of diagenetic overprint.
Lateral drainage was less important in providing
reservoir quality sandstones. Locations near paleohighs
commonly caused preservation of a condensed strati-
graphic sequence and distribution of poor quality sand-
CONCLUSIONS
stones. Observations of the structural and stratigraphic char-
Suspended load sedimentation, attributed to acteristics of several fields in the Cuyo basin indicate that
overbank flooding at extreme base level highstands, reverse faults that are prominent in the present structural
dominated the margins of synsedimentary depressions. configuration are genetically related to a system of linked
Reservoir quality is severely reduced or nonexistent extensional faults (see Gibbs, 1988) dominated by half-
along the margins of the half-grabens where the grabens and transfer zones. Mesozoic extensional
Mesozoic succession is thinner. This type of depositional faulting resulted in differential subsidence that accom-
pattern is recorded within the Potrerillos, Río Blanco, modated a 4000-m-thick clastic wedge. This tectonic
Barrancas, and Papagayos formations, the main produc- environment affected the distributions of both source
tive intervals in the area. rocks and reservoir rocks.
Cenozoic inversion of this preexisting extensional
Trapping fabric produced asymmetric structures that were
dissected by antithetic flank faults (back-thrusts) and that
Our observations of seismic sections indicate that the provided well-defined structural closures. The structural
oil-bearing traps are linked in variable degrees to the culminations in the productive trends show east- and
principal or secondary listric normal faults that west-vergent axial planes. Many seismic sections show
controlled depositional thickening. In particular, the that folds at shallow and intermediate depths are seated
original fabric comprising a suite of linked extensional above deeper faulted structures of the previous exten-
faults played a critical role in guiding stress during sional block architecture. Several anticlinal cores involve
inversion at a time when the faults acquired reverse thicker successions than those occurring along their
separation and most of the structural closures were flanks.
formed. During the latest Cenozoic, at the climax of Andean
Productive structures are thus related to the inversion orogenesis, the Mesozoic extensional structures were
processes that produced the structural culminations and comprehensively inverted and the structural traps
plunging noses, combined with basin blocks having developed to their present configurations. These anti-
favorable stratigraphy, such as reservoirs with access to clines and structural noses have the capacity to reservoir
source rocks. This occurs at Tupungato, Piedras Colo- oil volumes in the range of 200 million m3.
radas, and Chañares Herrados fields (Figure 5).
Closures at several fields appear to be related to the
transfer offsets of antithetic fault systems enhanced by
stratigraphic condensation and reservoir pinchouts along
the flanks of older half-grabens (e.g., Vizcacheras). Some
of the pools are associated with structural culminations
produced by transfer zones that connect grabens with
opposing polarities. In those positions, the traps coincide Acknowledgments The authors wish to acknowledge Astra
with wedges bounded by reverse faults (“pop-up” struc- C.A.P.S.A. and Enterra Oil Field Rental for release of data.
tures). Because these features developed by reactivation Raul Genovesi drafted the illustrations, and Fiona Shakespear
of structures that were already elevated, reservoir devel- helped with the English translation. Critical reviews by Miguel
opment was restricted to a few specific intervals Uliana and editorial work by Tony Tankard and Herman
(Barrancas, La Ventana, and Vacas Muertas fields). Welsink substantially improved the paper for publication.
Structural Inversion and Oil Occurrence in the Cuyo Basin of Argentina 367

REFERENCES CITED Nacional de Cuyo, Facultad de Ingeniería de Petróleos,


Mendoza, Serie 6 (1), p. 1–39.
Bally, A. W., 1984, Tectogenese et sismique réflexion: Societé Pombo, R., 1986, Areas en licitación Cuenca Cuyana: En
Géologique de France Bulletin 7, p. 279–285. Petróleo y Gas en Argentina, Buenos Aires, v. I, p. 151–163.
Biddle, K. T., and K. Rudolph, 1988, Early Tertiary structural Rolleri, E., and P. Criado Roque, 1968, La Cuenca triásica del
inversion in the Stord basin, Norwegian North Sea: Journal norte de Mendoza: Terceras Jornadas Geológicas
of the Geological Society, v. 145, p. 603–611. Argentinas, v. 1, p. 1–76.
Chebli, G., I. Labayén, G. Laffite, and M. R. Rosso, 1984, Rosso, M., I. Labayén, G. Laffite, and M. Arguijo, 1987, La
Materia orgánica, ambiente deposicional y evaluación generación de hidrocarburos en la Cuenca Cuyana:
oleogenética de la Cuenca Cuyana: Noveno Congreso Décimo Congreso Geológico Argentino, Actas II,
Geológico Argentino, Actas VII, p. 68–85. p. 267–270.
Criado Roqué, P., E. Rolleri, C. De Ferraris, I. Simonato, A. Turic, M., C. Fernández Garrasino, R. Pombo, J. L. Bianchi,
Suero, and T. Suero, 1960, Cuencas Sedimentarias de la and H. Di Benedetto, 1981, Cuencas Sedimentarias en la
Argentina: Boletin de Informacion Petroleras, Yacimientos Argentina: Comunicaciones: Yacimiento Petrolíferos
Petrolíferos Fiscales, v. 320, p. 62–95. Fiscales, Capacitación y Desarrollo, p. 14–18.
Gibbs, A., 1988, Balancing geoseismic cross-sections: Williams, G. D., C. M. Powell, and M. A. Cooper, 1989,
Canadian Society Petroleum Geology, Short Course, Geometry and kinematies of inversion tectonics, in M. A.
p. 1–88. Cooper and G. D. Williams, eds., Inversion tectonics:
Harrington, H. J., 1941, Investigaciones geológicas en las Geological Society Special Publication 44, p. 3–15.
Sierras de Villavicencio y Mal País: Dirección de Minería y
Geología, Boletín 49.
Jordan, T., and A. Ortíz, 1987, Tiempo de generación de
petróleo en Mendoza norte: Décimo Congreso Geológico
Argentino, Actas II, p. 271–276. Authors’ Mailing Addresses
Kokogian, D., and O. Mansilla, 1989, Análisis estratigráfico
secuencial de la Cuenca Cuyana, in G. Chebli and L. Daniel Dellapé
Spalletti, eds., Cuencas Sedimentarias Argentinas: Univer- Astra C.A.P.S.A.
sidad Nacional de Tucumán, serie Correlación Geológica, Tucumán 744
n. 6, p. 169–210. 1049 Buenos Aires
Legarreta, L., D. Kokogian, and D. Dellapé, 1993, Estruc- Argentina
turación terciaria de la Cuenca Cuyana: Cuanto de
inversión tectónica?: Asociación Geológica Argentina, Andrés Hegedus
Revista XLVII, p. 83–86.
Enterra Oil Field Rental
Mpodozis, C., and S. M. Kay, 1990, Provincias magmáticas
ácidas y evolución de Gondwana: Andes Chilenos Ruta 22 entre calles Caracas y Lima
(28–31˚ S): Revista Geológica de Chile, v. 17, p. 153–180. Colonia Valentina
Padula, E., 1972, Las Cuencas sedimentarias petrolíferas 8300 Neuquén
argentinas. Sus resultados: Serie G, 1, Universidad Argentina
Inversion of the Mesozoic Neuquén Rift in the
Malargüe Fold and Thrust Belt, Mendoza, Argentina

René Manceda Daniel Figueroa


Area Exploración Neuquén, YPF S.A. Proyectos Exploratorios. YPF S.A.
Neuquén, Argentina Buenos Aires, Argentina

Abstract

T he Malargüe fold and thrust belt formed by Mesozoic rift inversion during Tertiary compressional
orogeny. Mesozoic extension created the Neuquén basin in west-central Argentina and controlled most
structural styles and the geometry of the fold and thrust belt, which is characterized by basement-cored oppo-
sitely verging structures. Subsidence curves and palinspastically restored isopach maps of Mesozoic sedimen-
tary fill describe a complex pattern of asymmetric half-grabens bounded by major faults of opposite polarity
and accommodation zones related to a rift phase during Late Triassic–Early Jurassic time, as well as Middle
Jurassic–Early Cretaceous postrift regional subsidence. Balanced cross sections show the relationship between
preexisting extensional fabrics and contractional basement-involved thrusts and back-thrust structures that
generated the half-graben inversion. Overpressured shales and three evaporite levels favored formation of
duplexes, triangle zones, and detachment of the cover as a result of basement-involved shortening.

Resumen

L a Faja Fallada y Plegada de Malargüe resultó de la inversión tectónica del rift mesozoico, generada por
los movimientos compresivos terciarios. Este sistema extensional, que originó la Cuenca Neuquina en el
centro oeste de Argentina, controló en gran parte el estilo y la geometría de la faja fallada y plegada, caracteri-
zada por involucrar el basamento en el núcleo de estructuras con vergencias opuestas. El análisis del relleno
sedimentario del rift mesozoico mediante curvas de subsidencia y mapas isopáquicos palinspastizados,
permitió caracterizar depocentros relacionados con fallas maestras de polaridad cambiante, que se vinculan
mediante zonas de acomodación. Se determinó también una etapa rift durante el Triásico Superior-Jurásico
Inferior y una etapa de subsidencia termal durante el Jurásico Medio-Cretácico Inferior. La relación entre la
geometría inicial y la estructura actual de la faja fallada y plegada se muestra mediante secciones estructurales
balanceadas en las que el basamento se halla afectado por corrimientos y retrocorrimientos que inviertieron los
hemigrabenes. La presencia de niveles evaporíticos y pelitas euxínicas sobrepresionadas en cobertura favoreció
la formación de duplexes, zonas triangulares y pliegues de despegue.

INTRODUCTION The aim of this paper is to document the extensional


architecture of the Mesozoic basin and its control on the
The Malargüe fold and thrust belt is that part of the subsequent development of the Malargüe fold and thrust
Cordillera de los Andes between the Diamante and belt. Reconstruction is based on an analysis of the
Barrancas rivers (Figure 1). This fold and thrust belt is the Jurassic rift fill, palinspastically restored isopach maps,
climax of a long history of Phanerozoic geologic develop- and subsidence curves. Together, these elements describe
ment along the Pacific margin of South America, the extensional framework of the Malargüe basin tract.
including reactivation of old terrane boundaries to form Structural cross sections contrast this rift geometry with
Paleozoic and Mesozoic basins. the results of Tertiary compression and basin inversion.

Manceda, R., and D. Figueroa, 1995, Inversion of the Mesozoic Neuquén rift in the 369
Malargüe fold and thrust belt, Mendoza, Argentina, in A. J. Tankard, R. Suárez S., and
H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 369–382.
370 Manceda and Figueroa

Figure 1—Geologic map of the Malargüe tract of the Neuquén basin. (Adapted from Kozlowski, 1992.)
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 371

GEOLOGIC SETTING were 20 outcrop sections, mostly measured by YPF geol-


ogists. All surface and subsurface data were palinspasti-
The Malargüe fold and thrust belt is a Cenozoic defor- cally restored. The restored Jurassic isopach maps
mational system that resulted from the interaction of describe the original extensional geometry of the
Pacific margin subduction with preexisting extensional Neuquén basin. This interpretation was tested by subsi-
basin structures. The successive phases of the Andean dence curves using the Bond–Kominz method (1984).
orogeny deformed the sedimentary fill of the Neuquén Bond and Kominz include porosity and cementation
basin, as well as older rocks such as the Permian–Triassic changes in their calculation of the uncompacted
Choiyoi Group. The older units were essentially precur- thickness. We have derived porosities from well logs and
sors to the Neuquén successor basin, forming a basement thicknesses from detailed outcrop sections. Relative sea
to the Neuquén cover sequence (Figure 2). level changes and ages were inferred from Haq et al.
Most of the Argentine Mesozoic basins are attributed (1988). This rigorous approach avoids overestimation of
to extensional processes related to fragmentation of the uncompacted initial thicknesses (Figure 3).
Gondwana and the opening of the South Atlantic Ocean
(Uliana et al., 1989). There were three principal periods of
basin formation. (1) An early episode of Triassic rifting
was widespread, but was aborted without formation of EVOLUTION OF MALARGÜE BASINS
ocean basins. (2) Renewed extension in the Jurassic even- This paper focuses on the rift basins and their patterns
tually culminated in continental fragmentation. (3) The of subsidence in La Valenciana area (Figure 1), using
Cretaceous episode encompassed the transition to post- curves that discriminate between total and tectonic subsi-
rift regional subsidence, which is generally believed to dence (Figure 3). Each half-graben was analyzed inde-
reflect thermal decay. pendently.
The Neuquén basin records all three episodes, The Jurásico cycle spans Hettangian–earliest Kimmer-
although Jurassic extension by reactivation of older idgian time (Figure 2). Internally, it consists of intermit-
crustal fabrics was the most important (Ramos, 1989; tent rift sedimentation and an upper part attributed to
Uliana et al., 1989). Jurassic extension also records more thermal decay of the rift system. The tectonic subsidence
widespread sedimentation and volcanism, implying an curve (Figure 3) summarizes this history. Vigorous
expansion of the fault-controlled Triassic framework rifting is reflected in the coarse-grained Remoredo and
(Uliana et al., 1989). Reflection seismic sections and field lower El Freno formations. Fault-controlled subsidence
work in Neuquén (Gulisano, 1981; Gulisano and Pando, gradually diminished so that the upper El Freno and
1981) suggest that Jurassic extension resulted in a tract of Puesto Araya formations are distinctly finer grained.
tilted blocks. Seismic data show that the half-grabens During the late Toarcian, the rift-partitioned basin was
subsided along boundary faults of various orientations amalgamated into a broad epeiric sag. This new phase of
(Uliana et al., 1989). There have been numerous studies regional subsidence persisted until the Callovian when
of the Jurassic stratigraphy (summarized in Gulisano, there was a reversal in the trend of subsidence, as
1981; Legarreta and Gulisano, 1989; Legarreta and evidenced by the intra-Callovian unconformity (Dellapé
Uliana, 1991). Figure 2 shows the stratigraphy of the et al., 1979). Initially, limestones and evaporites were
Jurassic section. deposited on a stable platform. The end of the Oxfordian
Jurassic stratigraphy is complex because extension was marked by a major diastrophism event called the
created numerous depocenters that were subjected to Araucanian phase (Stipanicic and Rodrigo, 1970). An
different rates of subsidence and sediment supply, as intra-Malmic unconformity is related to development of
well as provenance. Fault-controlled subsidence was the adjacent magmatic arc. A relative fall of sea level is
conspicuous in Middle Triassic–Sinemurian time, but reflected in uplift and erosion of marginal areas and
subsequently succumbed to a more regional pattern of development of nonmarine depositional environments
subsidence as the various depocenters were amalga- (Riccardi and Gulisano, 1990). In post-Araucanian time
mated (Legarreta and Gulisano, 1989; Legarreta and (Kimmeridgian–Early Cretaceous), the basin resumed its
Uliana, 1991). Nevertheless, an irregular topography pattern of regional subsidence. The reversal of the subsi-
persisted, especially adjacent to the principal faults, dence trend in Albanian–Campanian time is attributed to
where it influenced the distribution of the Jurassic flexural deformation in response to thrust belt loading
megasequences. Although the early horst and graben (Ramos, 1988) or upper crustal emplacement of
topography of the Neuquén basin reflects extensional magmatic material (Legarreta and Uliana, 1991).
processes (Digregorio et al., 1984; Uliana et al., 1989), the Vigorous rift subsidence lasted about 25 m.y., and the
detailed structural framework is poorly understood. thermal subsidence phase another 70 m.y. Outcrop and
subsurface data show an apparent offset of rift and
postrift depocenters, a phenomenon possibly explained
METHODOLOGY by extensional detachment theory (see Kusznir et al.,
1987). Figure 3 shows a broadscale correlation of strati-
In an attempt to reconstruct the initial geometry of the graphic sequences with tectonic events related to styles
basin, we have analyzed 131 petroleum exploration wells of subsidence or periodic inversion. Uliana et al. (1989)
that were drilled through the Chacay-Lotena groups and show that depositional patterns were dominated by
locally encountered pre-Cuyo geology. Also integrated active normal faulting until the Toarcian, in agreement
372 Manceda and Figueroa

Figure 2—Stratigraphic column of the northwestern Neuquén basin. (Adapted from Groeber, 1946; Legarreta and
Gulisano, 1989.)
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 373

Figure 3—Subsidence curves of the La Valenciana half-


graben based on the Bond-Kominz back-stripping method
(Bond and Kominz, 1984). The tectonic subsidence curve
shows vigorous subsidence from 215 to 180 Ma. From 180
to 112 Ma, the subsidence gradually decreases (thermal
sedimentation).

with the tectonic subsidence curve. This rift to thermal


transition occurred in the Jurásico cycle of the lower
supersequence. Therefore, it does not fit the depositional
sequence boundaries as they are defined. However, a
broadscale subsidence model addressing the sequence Figure 4—Palinspastically restored isopach map of the
boundaries remains to be developed. complete Jurásico cycle, based on well and surface data.
Local areas of subsidence (e.g., La Valenciana-Atuel,
Palauco, and Chacay Melehue half-grabens) are separated
by intervening highs (e.g., Dedos-Silla, Bardas Blancas,
JURASSIC RIFT GEOMETRY and Río Grande valley). Locations of cross sections A–E
(in Figures 10–14) are shown. Contour interval is 100 m.
Palinspastically restored isopach maps of the Jurassic
succession (Figures 4, 5, 6, 7) show a suite of asymmetric
and discrete depocenters, a geometry we interpret as argillaceous or carbonate dominated (Leeder and
half-grabens. Figures 4, 5, and 6 show this half-graben Gawthorpe, 1987; Blair and Bilodeau, 1988). These char-
form expressed in various isopach groupings, for acteristics of basin geometry and sediment distribution
example, Jurassic succession, Cuyo Group, and pre-Cuyo allow us to interpret the Malargüe basins on the basis of
intervals. Cuyo deposition preserves a major component well control. (Subsequent Andean deformation makes
of rift subsidence. Higher in the succession, Chacay- the seismic expression of basin form even more tenuous.)
Lotena group isopachs (Figure 7) show a rearrangement Wedge-shaped sedimentary packages and shallowing
of depocenters, such as abandonment of the Palauco toward roll-overs (culminations) and accommodation
depocenter. These differences reflect rift-controlled depo- zones are observed.
sition of the lower Cuyo and postrift subsidence with a The isopach maps, subsidence curves, and strati-
strong regional signature for the upper Cuyo and graphic information together characterize the Jurassic
Chacay-Lotena interval. basin as follows:
The rates of subsidence along rift boundary faults
commonly exceed the rates of sedimentation supply, • There were separate phases of rift and postrift
resulting in underfilled basins and restricted marine or subsidence, each of which subsided intermittently.
lacustrine environments. In these basins, coarse deposits • Half-grabens are of various sizes, and isopach
stack close to the fault plain, while distal areas are more maps suggest changes in polarity of the principal
374 Manceda and Figueroa

Figure 5—Palinspastically restored isopach map of the Figure 6—Palinspastically restored isopach map of the
Cuyo Group, which mimics the previous structural relief. Cuyo Group and pre-Cuyo stratigraphy. These two
Contour interval is 100 m. intervals combined accentuate the structural geometry of
the basins. Contour interval is 100 m.

basin-forming faults. detachment model (see Wernicke, 1985; Kusznir et al.,


• The basins are separated by intervening highs. 1987). However, we have no direct evidence for such a
• There is a marked lateral variation in the distribu- detachment.
tion of sedimentary facies with euxinic facies The Atuel-Valenciana half-graben (Figures 4, 8, 9) has a
adjacent to boundary faults. westward-facing polarity, a Jurassic succession thickness
• Sedimentation started at different times in different greater than 2000 m, and an area of 9000 km2. This may,
basins, a characteristic also recorded by Groeber et in fact, be a composite feature with more than one half-
al. (1953) and Gulisano (1981). In the Atuel River graben; there may be one in the Atuel area and another
area, the Cuyo Group began in the Hettangian in the La Valenciana area, with a transfer zone between
(Riccardi et al., 1988). In contrast, in the Bardas them. However, the data are ambiguous. Cuyo sedimen-
Blancas area, sedimentation began during the late tary rocks are as old as Hettangian (Riccardi et al., 1988).
Toarcian (Gulisano, 1981), while in the Neuquén Deep water facies outcrop toward the east. These
province, the oldest sediments in the Cuyo Group deposits thin westward toward Portezuelo Ancho where
are Pliensbachian in age (Gulisano and Gutierrez they are no older than Sinemurian (Groeber et al., 1953).
Pleiming, 1984). The Palauco half-graben (Figures 4, 8, 9) is interpreted
from exploration wells. A thick clastic and volcanic
We believe that the apparent offsets of rift and postrift succession of Lias age (Papú and Zavattieri, 1989) is
depocenters and the along-strike variation in initiation of mapped. The position of the depocenter and the steepest
rift subsidence are best explained by an intracrustal slopes of the isopachs along the western border imply an
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 375

zones and transfer faults. Both trends affect the distribu-


tion and thickness of sedimentary facies. In the
Malargüe area, north-south trends are observed (e.g.,
Dedos-Silla area) with condensed sequences (Davidson
and Vicente, 1972). Legarreta and Kozlowski (1984)
observed similar trends in the Andico and Riograndico
cycles. We associate these sequences with roll-over
structures (Figures 4, 8).
The Bardas Blancas accommodation zone (Figures 4,
8, 9) is a composite feature consisting of a conjugate
divergent overlapping transfer zone between the Atuel-
Valenciana and Palauco half-grabens and a conjugate
convergent approaching transfer zone between the
Atuel-Valenciana and Chacay Melehue half-grabens
(Figures 8, 9) (see Scott and Rosendahl, 1989; Morley et
al., 1990). The Rio Grande valley high (Figures 4, 9)
(Kozlowski, 1992) straddles a synthetic collateral transfer
zone between the Chacay Melehue and Palauco troughs.

RELATIONSHIP OF TERTIARY
FOLDING TO JURASSIC EXTENSIONAL
STRUCTURES
Five balanced cross sections for the Malargüe fold and
thrust belt were prepared using surface data, reflection
seismic data, and exploration wells. The sections were
selected to show the present relationship between half-
grabens and Tertiary deformational relief (Figures 1, 4).

Atuel-Valenciana Cross Sections


Balanced cross sections A, B, and C from the Dia-
mante, Salado, and Malargüe rivers areas (Figures 10, 11,
12) show that the faults involving the basement in defor-
Figure 7—Palinspastically restored isopach map of the mation have planes dipping to the west, thus generating
Chacay-Lotena Group. Amalgamation of depocenters and foreland structures with steep to overturned beds in the
a regional pattern of subsidence are apparent. We attribute forward limb. These large-scale structures resulted from
this to postrift processes. Contour interval is 100 m. fault-propagation folding and fault-bend folding.
Deformation structures developed in the Jurassic
units west of the Diamante River (Figure 10), in the Los
east-facing polarity of the main fault. Small depocenters Blancos area (Figure 11), and in the La Valenciana area
occur toward eastern Palauco (Figure 14), with similar (Figure 12) display a vergence to the foreland, coin-
geometric characteristics. The thickness of the Palauco ciding with the initial rift geometry. South of the Atuel
half-graben fill exceeds 1200 m; its area is about 1800 km2. River, this shortening has resulted in triangle zones.
The Chacay Melehue half-graben (Figures 4, 9) is inter- These triangle zones involve Mendoza Group and
preted from a smaller data base. Nevertheless, Groeber et younger units as well, showing a displacement of the tip
al. (1953) and Gulisano and Gutierrez Pleiming (1984) line to the east and generating back-thrusts and passive
have documented great thicknesses of sedimentary roof duplexes. These are observed in the Salado River
successions in this area, while toward the east in the cross section (Figure 11) to the east of Cañada Ancha, in
Sierra Reyes, a transition from deep water to platform the Puesto Rojas–Cerro Mollar area (Figure 1) (Plos-
Cuyo facies occurs (Gulisano, 1981). We infer an zkiewicz, 1987), and in the Malargüe River cross section
eastward-facing polarity for the basin-forming listric (Figure 12). The surface expression of these triangle
normal fault. The depocenter continues toward the zones is a strongly deformed narrow belt of evaporites
south, where it is more than 2000 m thick and has a of the Huitrín Formation (Ploszkiewicz and Gorroño,
preserved area of 12,000 km2. 1988).
In extensional settings, basement highs are formed by Detachment folds are observed in the Salado River
structural relief of normal extensional fault systems, as cross section (Figure 11) between the Las Leñas Valley
well as the transverse components of accommodation and Los Molles.
376 Manceda and Figueroa

Figure 8—Isometric block


diagram of half-grabens
and highs based on
complete Jurásico cycle
data.

(below)
Figure 9—Structural rela-
tionships among half-
grabens and interpretation
of the basin-forming
extensional fault system.
Generalized cross
sections show changes in
structural style along
strike and changes in
polarity of the master
faults.
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 377

Figure 10—Cross section A, Diamante River area. The Atuel half-graben in the west is inverted, generating a complex duplex
system. The structure in the Diamante River area involves basement and reflects the southward plunge of the Cordón del
Carrizalito. See Figure 4 for location.

Figure 11—Cross section B, Salado River area. Note the detachment folding in front of the Las Leñas anticline and the
triangle zones east of the Los Blancos anticline. (Cañada Ancha zone was adapted from Condat et al., 1989.) See Figure 4 for
location.
378 Manceda and Figueroa

Figure 12—Cross section C, Malargüe River area. La Valenciana anticline is formed by inversion of the Jurassic depocenter,
producing a triangle zone at its edge. The presence of back-thrusts in the Tricolor hill are believed to be subordinate to the
main deformation; it verges toward the foreland. See Figure 4 for location.

Accommodation Zone Cross Section quently, the hanging wall of the back-thrust involves
basement. Hayward and Graham (1989) have described
Similar characteristics of deformation affected the these characteristics as a shortcut fault. Toward the east,
basement-cored accommodation zones with a duplex the available data suggest that the inverted faults have
verging toward the foreland in Bardas Blancas (Manceda rooted in the synrift succession without generating
et al., 1990, 1992) and fault-bend folding in La Batra, shortcuts (see Suppe, 1986). Asymmetry toward the
where it is associated with a southward plunge of the foreland forms surface topography in the Sierra de
Malargüe anticline (cross section D, Figure 13). Palauco, Ranquilco-Del Petiso, and Cerro Fortunoso
In contrast, the Portezuelo del Viento structure and La areas.
Zeta and La Guanaca faults (Figure 13) as well as the
subsurface Los Cerrillos structure verge in the opposite Interpretation
direction. These structures are not associated with
inversion tectonics, nor are they related to a Jurassic The cross sections show that the Atuel-Valenciana and
depocenter. Instead, they appear to reflect adjustment to Palauco half-grabens were inverted by Tertiary compres-
fold and thrust belt deformation. sional tectonics, forming large-scale structures. Neverthe-
The sedimentary cover, which drapes a thin Jurassic less, there are other anticlinal structures in the foldbelt,
succession along the accommodation sidewall, has such as Sierra Azul-Bardas Blancas, Sierra de Cara Cura,
developed duplexes similar to the belt found between Sierra de Reyes, and Bloque Dedos-Silla (the Dorsal del
the Campanario Mountain and Portezuelo del Viento Tordillo of Davidson and Vicente, 1972, and Legarreta
and the triangle zone of Bardas Blancas (Plozkiewicz, and Gulisano, 1989), that do not correspond to the
1987; Manceda et al., 1990). thickest Jurassic synrift deposits. These are attributed to
structural inversion of normal faults that formed the
Palauco Rift Cross Section internal structure of the half-grabens. However, these
antithetic secondary faults are difficult to map because
Structural inversion is conspicuous in the Palauco the database is not dense enough to record their subtle
half-graben where Tertiary compressional tectonics have form.
reactivated the basin-forming listric normal faults in a Although tenuous, the large-scale structure associated
reverse sense. Both the Jurassic fill and the basement with the Cordillera del Viento is attributed to structural
have been compressed and elevated (cross section E, inversion of the Chacay Melehue half-graben. This has
Figure 14). The asymmetry of the structures toward the generated a back-thrust toward the west that explains
west is based on abundant well control. A significant the abrupt relationship between basement and the
characteristic of the Palauco structure is that only the synorogenic prism. To the east, the Jurassic succession
low-angle listric normal fault has been reactivated, while lies unconformably on the basement which at the surface
the high-angle component has not been affected. Conse- dips slightly to the east.
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 379

Figure 13—Cross section D, Rio Grande area. As cross section C, the principal deformation is toward the foreland; there are
also subordinate back-thrusts. The thickness of the Jurassic succession has decreased adjacent to the accommodation
sidewall. Contraction of the basement is absorbed in duplexes that also affect the sedimentary cover to the west. Toward the
east, deformation is expressed in triangle zones (Bardas Blancas). See Figure 4 for location.

The Rio Grande high forms the Sierra Azul anticline records the initiation of Malargüe deformation. Toward
and an associated structural depression to the east that is the west, it is probable that older ages will be found.
expressed as the Río Grande valley. This depression is The average shortening derived from balanced cross
confined to the west by the thrust structure and to the sections is 38 km, suggesting an average rate of deforma-
east by back-thrusts of the Palauco anticline. This depres- tion of 0.7–0.9 mm/yr.
sion has a complex structure within the cover where
opposing vergences occurred in the late Miocene.
The sudden changes in vergence and structural styles CONCLUSIONS
that were controlled by the Jurassic rift infrastructure are
clearly reflected in the present north-south structural The Malargüe fold and thrust belt has been explored
trends. This is also true for the transfer shortening for many years. Numerous hydrocarbon discoveries
between the Malargüe structure and the inverted have been made in deformed structures that had an
depocenter in Palauco, where vergences are opposite, earlier history of subsidence along extensional fault
and between the Bardas Blancas structure and the systems. This exploration database has been used to
inverted depocenter of La Valenciana, where vergences reconstruct the history of basin subsidence and subse-
are similar. quent inversion. The principal results of this study are as
follows:
Timing and Rate of Deformation
• Jurassic sedimentation initially occurred in half-
The Tertiary synorogenic prism is separated from grabens that had regular along-strike changes in
older basin fills and basement by angular unconformi- polarity. Such changes in polarity, timing, and
ties. The earliest synorogenic deposits in the external part structural styles imply offsetting accommodation
of the foldbelt are dated as late Eocene on the basis of zones (see Gibbs, 1984; Lister et al., 1986).
Deseadense age mammals (Gorroño et al., 1979), about • Subsidence is attributed to rift and postrift
39 Ma. Based on stratigraphic considerations, Uliana and processes, each with at least two episodes of
Dellapé (1981) and Legarreta and Gulisano (1989) have enhanced subsidence.
suggested an early Eocene age, nearly 55 Ma, but there is • The principal basin-forming faults of the half-
no paleontologic confirmation. This angular unconfor- grabens, and even the secondary internal fabrics of
mity between the synorogenic prism and the Malargüe these basins, suffered varying degrees of reactiva-
Group, which is the youngest sedimentary unit of the tion and inversion during the Tertiary Andean
Riográndico cycle (Legarreta and Gulisano, 1989), orogeny.
380 Manceda and Figueroa

Figure 14—Cross section E, Sierra de Palauco area. Coincident with the initial extensional geometry, compression has
resulted in a change in the sense of deformational vergence. Inversion of the Palauco half-graben toward the west is
observed. See Figure 4 for location.

• Basement involvement and the opposing directions Bond, G., and M. Kominz, 1984, Construction of tectonic
of vergence were controlled by the initial architec- subsidence curves for the early Paleozoic miogeocline,
ture of the basin. southern Canadian Rocky Mountains: implications for
• Shortening derived from balanced cross sections subsidence mechanisms, age of breakup, and crustal
thinning: GSA Bulletin, v. 95, p. 155–173.
averaged 32%, suggesting a mean rate of deforma-
Condat, P., E. Kozlowski, E. C. Cruz, and R. Manceda, 1989,
tion of 0.7–0.9 mm/yr. Estructura del anticlinal Cañada Ancha, Mendoza: Boletín
de Informaciones Petroleras, Tercera Epoca, no. 19, p. 2–14.
Davidson, M. J., and J. C. Vicente, 1972, Características paleo-
Acknowledgments We wish to thank YPF S.A. for geográficas y estructurales del área fronteriza de las
allowing publication of this paper and our colleagues in the nacientes del Teno (Chile) y Santa Elena (Argentina),
Neuquén Exploration Area and Development Projects Area in Cordillera Principal, 35° a 35° 15’ de latitud Sur: Quinto
Buenos Aires for helpful discussions. V. A. Ramos, E. Congreso Geológico Argentino, v. 5, p. 11–55.
Dellapé, D. A., C. Mombrú, G. Pando, A. C. Riccardi, M. A.
Kozlowski, H. Welsink, and especially A. Tankard provided
Uliana, and G. E. Westermann, 1979, Edad y correlación de
careful review of this manuscript, and their comments served la Formación Tábanos en Chacay Melehue y otras locali-
to greatly improve it. dades de Neuquén y Mendoza. Con consideraciones sobre
la distribución y significado de las sedimentitas lotenianas:
Obra Centenario del Museo de La Plata 5, p. 81–105.
Digregorio, R. E., C. A. Gulisano, A. R. Gutierrez Pleimling,
REFERENCES CITED and S. A. Minitti, 1984, Esquema de la evolución geod-
Blair, T., and W. Bilodeau, 1988, Development of tectonic inámica de la Cuenca Neuquina y sus implicancias paleo-
cyclothems in rift, pull-apart, and foreland basins: sedi- geográficas: Noveno Congreso Geológico Argentino, Actas
mentary response to episodic tectonism: Geology v. 16, 2, p. 147–162.
p. 517–520. Gibbs, A. D., 1984, Structural evolution of extensional basin
Inversion of the Mesozoic Neuquén Rift, Malargüe Fold and Thrust Belt, Mendoza, Argentina 381

margins: Journal of the Geological Society of London, v. changes at active plate margins: processes and products:
141, p. 609–620. International Association of Sedimentologists Special
Gorroño, R., R. Pascual, and R. Pombo, 1979, Hallazgo de Publication 12, p. 429–450.
mamíferos eógenos en el sur de Mendoza, su implicancia Lister, G. S., M. A. Etheridge, and P. A. Symonds, 1986,
en la datación de los “Rodados Lustrosos” y del primer Detachment faulting and the evolution of passive conti-
episodio orogénico del Terciario en esa región: Séptimo nental margins: Geology, v. 14, p. 246–250.
Congreso Geológico Argentino, Actas 2, p. 475–487. Manceda, R., E. Kozlowski, C. Cruz, and P. Condat, 1990,
Groeber, P., 1946, Observaciones geológicas a lo largo del Secuencia de techo pasiva estructurada, Bardas Blancas,
meridiano 70. Hoja Chos Malal: Sociedad Geológica Provincia de Mendoza, Argentina: Décimo Congreso
Argentina Revista, v. 1, p. 117–208. Geológico Argentino, Actas 2, p. 27–30.
Groeber, P., P. Stipanicic, and A. Mingramm, 1953, Jurásico: Manceda, R., E. Bolatti, and R. Manoni, 1992, Modelo estruc-
Geografía de la República Argentina, Sociedad Argentina tural para la zona de Bardas Blancas, Malargüe, Provincia
de Estudios Geográficos, Tomo 2, p. 143–347. de Mendoza: Boletín de Informaciones Petroleras, Tercera
Gulisano, C. A., 1981, El Ciclo Cuyano en el norte del Epoca, no. 31, p. 92–103.
Neuquén y sur de Mendoza: Octavo Congreso Geológico Morley C. K., R. A. Nelson, T. L. Patton, and S. G. Munn, 1990,
Argentino, Actas 3, p. 579–592. Transfer zones in the east African rift system and their
Gulisano, C. A., and A. R. Gutierrez Pleiming, 1984, Esquema relevance to hydrocarbon exploration in rifts: AAPG
estratigráfico de la secuencia jurásica del oeste de la Bulletin, v. 74, p. 1234–1253.
Provincia del Neuquén: Noveno Congreso Geológico Papú, H., and A. M. Zavattieri, 1989, Estudio palinológico de
Argentino, Actas 1, p. 236–259. las coronas de los sondeos YPF. Md. NRCo. x-8 (Ranquil-
Gulisano, C. A., and G. Pando, 1981, Estratigrafía y facies de co) e YPF. Md. NCL. x-4 (Cajon de Letelier): Centro
los depósitos Jurásicos entre Piedra del Aguila y Sañicó, Regional de Investigaciones Científicas y Técnicas,
Departamento de Collon Curá. Prov. de Neuquén: Octavo Mendoza, internal report, 4 p.
Congreso Geológico Argentino, Actas 3, p. 553–557. Ploszkiewicz, V., 1987, Las zonas triangulares de la faja
Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic and fallada y plegada de la cuenca neuquina, Argentina:
Cenozoic chronostratigraphy and cycles of sea-level Décimo Congreso Geológico Argentino, Actas 1,
change, in C. K. Wilgus, B. S. Hastings, Ch. G. St. C. p. 177–180.
Kendall, H. W. Posamentier, Ch. A. Ross and J. C. Van Ploszkiewicz, V., and R. Gorroño, 1988, Tectónica de
Wagoner, eds., Sea-level changes: an integrated approach: inyección salina en la faja fallada y plegada del sur de
SEPM Special Publication 42, p. 71–108. Mendoza: Boletín de Informaciones Petroleras, Tercera
Hayward, J., and R. Graham, 1989, Some geometrical charac- Epoca, no. 14, p. 29–34.
teristics of inversion, in M. A. Cooper and G. D. Williams, Ramos V. A., 1988, The tectonics of the central Andes: 30° to
eds., Inversion tectonics: The Geological Society Special 33° S latitude, in S. Clark and D. Burchfiel, eds., Processes
Publication 44, p. 17–39 in continental lithospheric deformation: GSA Special Paper
Kozlowski, E., 1992, Structural geology of the NW Neuquina 218, p. 31–54.
basin, Argentina: Cuarto Simposio Bolivariano “Explo- Ramos V. A., 1989, The birth of southern South America:
ración Petrolera en las Cuencas Subandinas,” Tomo 1, American Scientist, v. 77, p. 444–450.
p. 1–10. Riccardi, A. C., and C. A. Gulisano, 1990, Unidades limitadas
Kusznir, N. J., G. D. Karner, and S. Egan, 1987, Geometric, por discontinuidades: su aplicación al Jurásico andino:
thermal and isostatic consequences of detachments in Asociacion Geologica Argentina Revista, v. 45, p. 346–364.
continental lithosphere extension and basin formation, in Riccardi, A. C., C. E. Damborenea, M. O. Manceñido, and S. C.
C. Beaumont and A. J. Tankard, eds., Sedimentary basins Ballent, 1988, Hettangiano y Sinemuriano marinos en
and basin-forming mechanisms: Canadian Society of Argentina: Quinto Congreso Geológico Chileno, Actas 2,
Petroleum Geologists, Memoir 12, p. 185–203. p. C359–C373.
Leeder, M. R., and L. Gawthorpe, 1987, Sedimentary models Scott, D. L., and B. R. Rosendahl, 1989, North Viking graben:
for extensional tilt-block–half-graben basins, in M. P. an east African perspective: AAPG Bulletin, v. 73,
Coward, J. F. Dewey, and P. L. Hancook, eds., Continental p. 155–165.
extensional tectonics: The Geological Society Special Publi- Stipanicic, P., and F. Rodrigo, 1970, El diastrofismo jurásico en
cation 28, p. 139–152. Argentina y Chile: Cuartas Jornadas Geológicas
Legarreta L., 1991, Evolution of a Callovian–Oxfordian Argentinas, Actas 2, p. 353–368
carbonate margin in the Neuquén basin of west-central Suppe, J., 1986, Reactivated normal faults in the western
Argentina: facies architecture, depositional sequences, and Taiwan fold-and-thrust belt: Geologic Society of China,
global sea-level changes: Sedimentary Geology, v. 70, Memoir 7, p. 187–200.
p. 209–240. Uliana, M., and D. A. Dellapé, 1981, Estratigrafía y evolución
Legarreta, L., and C. A. Gulisano, 1989, Análisis estratigráfico paleoambiental de la sucesión maastrichtiana–eoterciaria
secuencial de la cuenca neuquina (Triásico del engolfamiento neuquino (Patagonia septentrional):
superior–Terciario inferior) Argentina, in G. Chebli and L. Octavo Congreso Geológico Argentino, Actas 3,
Spalletti, eds., Universidad Nacional de Tucumán, p. 673–671.
p. 221–243. Uliana, M. A., K. T. Biddle, and J. Cerdán, 1989, Mesozoic
Legarreta, L., and E. Kozlowski, 1984, Secciones condensadas extension and the formation of Argentine basins, in A. J.
del Jurásico–Cretácico de los Andes del sur de Mendoza: Tankard and H. R. Balkwill, eds., Extensional tectonics and
estratigrafía y significado tectosedimentario: Noveno stratigraphy of the north Atlantic margins: AAPG Memoir
Congreso Geológico Argentino, Actas 1, p. 286–297. 46, p. 599–614.
Legarreta L., and M. Uliana, 1991, Jurassic–Cretaceous marine Wernicke B., 1985, Uniform-sense normal simple shear of the
oscillations and geometry of back-arc basin fill, central continental lithosphere: Canadian Journal of Earth Science,
Argentine Andes, in D. I. M. Macdonald, ed., Sea-level v. 22, p. 108–205.
382 Manceda and Figueroa

Authors’ Mailing Addresses


René Manceda
Area Exploración Neuquén, YPF S.A.
Plaza Huincul, Neuquén
Argentina

Daniel Figueroa
Proyectos Exploratorios, YPF S.A.
Av. R. Sáenz Peña 777
1364 Buenos Aires
Argentina
Tectonic Evolution and Paleogeography of the
Neuquén Basin, Argentina
G. D. Vergani A. J. Tankard
Perez Companc Tankard Enterprises
Neuquén, Argentina Calgary, Alberta, Canada

H. J. Belotti
H. J. Welsink
Perez Companc
Neuquén, Argentina

Abstract

T he tectonic evolution of the Neuquén basin spans about 220 m.y. of Mesozoic–Cenozoic subsidence.
Initial rifting in the Triassic was driven by extensional collapse of the Permian–Triassic orogen. This
period of extension was accommodated by inherited structural inhomogeneities and a southwest-oriented
extensional stress field. From the Aalenian onward, fault-controlled subsidence was replaced by regional subsi-
dence. Several episodes of structural inversion modified the shape of the depocenter and rejuvenated fringing
sedimentary source areas. The most significant inversion occurred in the late Oxfordian–earliest Kimmeridgian
when the Dorsal de Huincul was formed. This Late Jurassic diastrophism marks a fundamental reorganization
of extensional stress fields related to fragmentation of southwestern Gondwana and the Atlantic opening. Late
Jurassic–Cretaceous extension was northwest directed. This history of tectonic evolution is reflected in a
complex structural framework, at least two major hydrocarbon source rock intervals, and numerous reservoir
zones.

Resumen

L a evolución tectónica de la cuenca Neuquina, desarrollada a través del Mesozoico y Cenozoico, abarca
unos 220 m.a. El inicio del rift en el Triásico tardío fue provocado por un colapso extensional del
orógeno permo-triásico. Este período de extensión fue motivado por un campo de esfuerzo extensional de
orientación sudoeste-noreste, e influenciado por inhomogeneidades estructurales inherentes al sustrato. A
partir del Aaleniano la subsidencia controlada por fallas fue reemplazada por subsidencia de carácter regional.
Varios episodios de inversión tectónica modificaron la forma del depocentro y rejuvenecieron las áreas de
aporte sedimentario. La inversión mas significante ocurrió en el Oxfordiano tardío–Kimmeridgiano temprano.
cuando se formó la Dorsal de Huincul. Este diastrofismo del Jurásico tardío, marca una reorganización funda-
mental de los campos de esfuerzo extensionales, relacionados con la fragmentación del sudoeste del
Gondwana y la apertura atlántica. La extensión acaecida durante el Jurásico–Cretácico tuvo orientación
noroeste-sudeste. Esta historia de evolución tectónica se refleja en un marco estructural complejo con al menos
dos intervalos de roca generadora de hidrocarburos y numerosas rocas reservorio.

INTRODUCTION extensional subsidence resulted in several source rock


intervals and Jurassic and Cretaceous reservoirs. The
The Neuquén basin of west-central Argentina is Neuquén accounted for about 44% of Argentinian oil
surrounded by the North Patagonian massif, the Sierra production in 1992 (Argentinian Petroleum Institute,
Pintada, and the Andean cordillera (Figure 1). It is a 1993). More than 100,000 km of reflection seismic have
triangular-shaped basin covering more than 160,000 km2 been acquired and over 1300 exploration wells drilled
that contains a Mesozoic–Cenozoic sedimentary succes- since exploration began in 1904 (Uliana and Legarreta,
sion (Figure 2) at least 7 km thick. A protracted period of 1993).

Vergani, G. D., A. J. Tankard, H. J. Belotti, and H. J. Welsink, 1995, Tectonic evolution and 383
paleogeography of the Neuquén basin, Argentina, in A. J. Tankard, R. Suárez S., and H.
J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 383–402.
384 Vergani et al.

Digregorio and Uliana, 1980; Gulisano et al., 1984a,b;


Mitchum and Uliana, 1985; Legarreta and Gulisano,
1989; Legarreta and Uliana, 1991; Uliana and Legarreta,
1993). Recent studies have compared the Jurassic–Creta-
ceous succession to global sea level charts (compare
Legarreta and Uliana, 1991, with Haq et al., 1987). The
aim of the present paper is to address the structural
framework of the Neuquén basin and the way its
evolution affected sedimentation.

GEOLOGIC SETTING
Patagonia preserves a long history of Phanerozoic
basin formation. In Late Permian–Early Triassic time, the
Paleozoic basins were subjected to an episode of wide-
spread compressive inversion that was relatively
selective. Conspicuous deformation belts include the
Sierras Australes, Sierra Grande, Precordillera, Sierra
Pintada, and South Patagonia (see Cobbold et al., 1986).
This period of mountain building was a precursor to
intermediate and acid magmatism of Permian–Triassic
age (Rapela and Kay, 1988; Kay et al., 1989; Uliana et al.,
1989). The Choiyoi Group (Figure 2) is a rhyolite-
ignimbrite suite of enormous proportions that is closely
associated with extensional tectonics and molasse depo-
sition. Uliana and Legarreta (1993) attribute these associ-
ations to extensional collapse.
The Neuquén basin records at least 220 m.y. of basin
subsidence. The preserved basin fill consists of an Upper
Triassic–Cenozoic succession that is at least 7000 m thick.
It is punctuated by several unconformities that reflect
intermittent subsidence, as well as several episodes of
structural inversion (Figure 2). The most conspicuous
inversion structure is the Huincul dorsal, or arch, of early
Figure 1—Location of Neuquén basin in west-central Kimmeridgian age. Structural inversion was an
Argentina, showing its relationship to the Sierra Pintada, important component of hydrocarbon trap formation.
North Patagonian massif, and the younger Andean
Cordillera to the west.
There are six principal stages of basin evolution. (1)
Norian–Sinemurian extension was accompanied by
molasse deposition. (2) Renewed extension in the Pliens-
The tectonic and stratigraphic evolution of the bachian–Toarcian involved extensional faults that parti-
Neuquén basin began during the Late Triassic and Early tioned the basin into several fault-controlled depocen-
Jurassic as a result of extensional collapse of the ters. This was also the time of Los Molles source rock
Permian–Triassic orogenic belt (see Dewey, 1988). deposition. (3) Aalenian–Oxfordian postrift regional
However, the basin has an even longer ancestry (Tankard subsidence resulted in amalgamation of the previous rift
et al., 1995). The Mesozoic Neuquén basin has a complex depocenters. (4) The latest Oxfordian and early
structural framework that was affected first by northeast- Kimmeridgian Araucanian diastrophism and inversion
oriented extension in the Triassic and Early Jurassic and marks a fundamental reorganization of extensional stress
then by northwest-directed extension related to Late fields. (5) Renewed subsidence in the Late Jurassic–Early
Jurassic–Cretaceous fragmentation of western Gondwana Cretaceous was initiated by extensional processes.
and the Atlantic opening. The most conspicuous structure Several unconformities record intermittent subsidence of
is the Huincul dorsal, or arch. It is attributed to Kimmerid- an epeiric basin. The Vaca Muerta is a prolific source
gian inversion of an earlier rift depoaxis; this deformation rock interval. (6) In the Late Cretaceous–Tertiary, the
is known as the Araucanian event. The importance of the Neuquén basin was subjected to progressive growth of
Huincul inversion is that it marked the reorganization of the Andean fold and thrust belt and reactivation of older
the Mesozoic stress fields and coincided with the initia- extensional faults in a reverse sense.
tion of rifting along the eastern seaboard of Argentina The principal tectonic elements comprising the
and the conjugate margin of South Africa. Neuquén basin are summarized in Figure 3. The
Numerous studies have addressed the stratigraphic essential ingredients that form this tectonic architecture
evolution of the Neuquén basin (e.g., Marchese, 1971; include Triassic–Jurassic extensional processes, multiple
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 385

Figure 2—Tectonostratigraphic column for the Neuquén basin showing major basin-forming stages. The unconformity-
bounded stratigraphy is based on seismic stratigraphy, well control, proprietary exploration reports, and literature. Los
Molles and Vaca Muerta are the principal source rock intervals. The relative sea level curve is based on Haq et al. (1987) and
adjusted to the DNAG (Decade of North American Geology) time scale. A–G, seismic markers.
386 Vergani et al.

Figure 3—Maps showing (A) principal tectonic framework of the Neuquén basin and (B) associated oil and gas fields and
locations of seismic lines and cross sections.

episodes of Mesozoic and Cenozoic inversion, and devel- Structural Styles


opment of the Andean fold and thrust belt and late
Tertiary foreland basin. This structural framework A series of superimposed structural styles affected the
controls the distribution of oil and gas fields. Neuquén basin throughout its evolution (Uliana and
Legarreta, 1993). The principal fault trends are north-
south, ENE-WSW, and northwest-southeast (Figure 4A).
From the Late Triassic through Early Jurassic, a system of
TRIASSIC–JURASSIC SUBSIDENCE extensional faults developed that formed the initial
Neuquén basin. Outcrop and well data suggest that
Early Rifting during this time an extensional system was active and
controlled accumulation of the pre-Cuyo and lower
The earliest episode of rifting is of Late Triassic–Sine- Cuyo Group sediments (Figures 4B, C).
murian age (Figure 2). It is characterized by basement- In the southern part of the basin, northeast-southwest
involved normal extensional faults that bound a suite of and ENE-WSW extensional faults are subparallel to the
discrete half-grabens (Figures 4A, B). These basins Limay River. The various depocenters are believed to
contain thick accumulations of coarse-grained sediments have been linked by a transfer fault system (Figures 4A,
that were deposited in continental environments, B, C). In the Dorsal de Huincul area and toward the
volcanics, and volcaniclastic materials. In the southwest, southern margin of the basin, various fault blocks plunge
these rocks crop out in the Chacaico and Sañico ranges northward along a normal fault system that is intersected
and in the Piedra del Aguila region (Digregorio and by en echelon transfer faults. This is observed in the El
Uliana, 1980; Gulisano and Pando, 1981). Their broader Sauce, Barda Colorada–Plaza Huincul, and Challacó
distribution (Figure 4B) is shown by reflection seismic areas (Figure 4A). In the eastern half of the basin, subsi-
data and exploration wells. Although a variety of formal dence was controlled by a northwest-southeast oriented
names have been assigned to these rocks, the less formal extensional fault system; examples include the Bajada
grouping as pre-Cuyo distinguishes these continental Vidal, Entre Lomas, Gobernador Ayala, and Rio
deposits from the succeeding marine deposits of the Colorado faults (Figure 4A). This fault system marks the
Cuyo Group (Figure 2). These pre-Cuyo strata underlie break between the platform and inner basin.
some of the deepest parts of the Neuquén basin. For this These main regional features are shown in seismic
reason they are the least studied and generally of sections of Bajada Vidal and Entre Lomas (Figures 5, 6).
unknown thickness. The Bajada Vidal fault dips westward toward the basin
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 387

Figure 4—Structural and paleogeographic framework of early Neuquén basin. (A) Extensional architecture of Late
Triassic–Early Jurassic rift phase. Abbreviations: A, Añelo; BC, Barda Colorada; BV, Bajada Vidal; CC, Cara Cura range;
Ch, Chacaico; CL, Catan Lil; CSN, Chihuido de la Sierra Negra; Co, Challacó; CV, Cordillera del Viento; DCh, Dorsal de los
Chihuidos; EL, Entre Lomas; ES, El Sauce; EV, Estancia Vieja; GA, Gobernador Ayala; Ha, Huantraico; PA, Piedra del
Aguila; PH, Plaza Huincul; RC, Río Colorado; RN, Rio Negro; Sa, Sañico; SR, Reyes range; TCh, Tres Chorros; VM, Vaca
Muerta. (B) Early rift paleogeographic reconstruction. (C) Lower Cuyo Group late rift paleogeography showing isolated rift
depocenters. (D) Upper Cuyo Group postrift phase in which previous fault-controlled depocenters were amalgamated in a
broad downwarp.
388 Vergani et al.

Figure 5—Seismic section showing Bajada Vidal trough and boundary fault, as well as the transition from rift subsidence to
the postrift cover sequence. Vertical exaggeration is 4.2:1 at 1.4 sec (two-way time). See Figure 3 (line D) for location.

and has generated a narrow half-graben along strike. basin by a horst that coincides with the Dorsal de los
Well and seismic data suggest that over 1500 m of synrift Chihuidos (Marchese, 1971).
sediments of the pre-Cuyo and lower Cuyo groups accu- In the western part of the basin, a north-south exten-
mulated in the half-graben (Figures 4B, C). The Bajada sional fault system controlled deposition of the Cuyo
Vidal fault is about 70 km long and appears to connect Group sediments. An important structural lineament
with the Estancia Vieja fault (Figure 4A). Toward the marks a depositional axis, from the Cordillera del Viento
north, the Entre Lomas fault is offset from the Bajada southward along Tres Chorros, Chacaico, Catan Lil, and
Vidal fault and is of opposite polarity (Figures 4A, 6). As Sañico. To the north, this trend extends into the Mendoza
a result, this fault has generated a narrow half-graben. province (Manceda and Figueroa, 1993). From the Early
These narrow, deep fault-bounded Entre Lomas and Jurassic onward, evolution of the Neuquén basin was
Bajada Vidal troughs suggest a component of transten- characterized by regional subsidence interrupted by
sional subsidence. periodic inversion.
The faults in the Gobernador Ayala and Río Colorado
areas merge westward with the north-south extensional Lower Jurassic Rift Stratigraphy
system (Figure 4A). The Chihuido de la Sierra Negra
faults belong to a basement high that limited the lower Transition to the Cuyo Group is marked by a lower
Cuyo Group sediments to the north (Figures 4A, C). Pliensbachian unconformity and deposits representative
Cuyo deposition was also constrained in the center of the of marine paleoenvironments. These lower Cuyo strata
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 389

Figure 6—Seismic section showing Entre Lomas fault-controlled trough and transition to postrift subsidence. The narrow,
deep trough suggests transtensional subsidence, possibly related to a Riedel shear. Vertical exaggeration is 3.2:1 at 1 sec
(two-way time). See Figure 3 (line E) for location.
390 Vergani et al.

Figure 7—Isopach distributions reflecting (A) northeast-oriented extensional regimes in the Early–Middle Jurassic and
(B) northwest-directed extension in the Late Jurassic–Early Cretaceous. Isopachs are based on seismic and well control and
integrated proprietary reports. (Courtesy of YPF and Perez Companc.)

have been dated as Pliensbachian–Toarcian (Gulisano Jurassic Postrift Stratigraphy


and Pando, 1981; Hinterwimmer and Jauregui, 1984;
Legarreta and Gulisano, 1989), although deposition in From late Toarcian time onward, fault-controlled
the Mendoza province started in the Hettangian subsidence gradually decreased over most of the
(Riccardi et al., 1990). The distribution and thickness of Neuquén basin. Nevertheless, local structural trends
these lower Cuyo rocks imitate the pre-Cuyo fault remained active even through Punta Rosada time, such
system, suggesting that extensional faulting was an as the Entre Lomas structure along the northern edge of
important control of sedimentation during the Early the basin. The isolated rift depocenters of earlier Jurassic
Jurassic. The thickest accumulations of marine strata are time (Figure 4C) were gradually amalgamated into a
confined to depocenters bounded by substantial struc- single, broad basin in which the previous interbasin
tural relief and marginal horst blocks (Figure 4C). This highs lost their identities (Figure 4D). Figure 7A shows
sequence is dominated by the argillaceous Los Molles the isopach geometry of this postrift epeiric basin. A
Formation, consisting of basinal shale, turbidites, and depositional axis persisted along the Piedra del
lenticular sandstone (Gulisano et al., 1984b; Hinter- Aguila–Barda Colorada–Chacaico trend (Figure 7A).
wimmer and Jauregui, 1984; Carbone, 1988). Geochem- However, basement topography was still expressed in
ical characterization shows that the Los Molles was the northwest-trending Dorsal de los Chihuidos. Deposi-
deposited in a variety of depocenters spanning saline to tional thickening also occurred along northwest-oriented
lacustrine settings (Zumberge, 1993). The Los Molles and Entre Lomas and Bajada Vidal depocenters (Figures 4A,
the coarse Lajas facies form an upward-coarsening depo- 7A). Figures 5 and 6 illustrate the seismic expression of
sitional sequence. In the Entre Lomas area, there is the rift-postrift transition along the northern ramp of the
evidence that Punta Rosada (Lajas equivalent) sedimen- basin. The Bajada Vidal trough formed a narrow but
tation was locally fault controlled. deep rift, possibly reflecting transtensional processes.
Figure 4C shows that lower Cuyo deposition was The postrift Cuyo Group forms a basinward-thickening
largely confined by structural depocenters of Late cover sequence that was locally affected by fault activity.
Triassic–Early Jurassic origin. The structural framework These characteristics of basement geometry, including
suggests that extension used northeast-oriented accom- northwest-trending basement topography shown on the
modation zones. This interpretation supports the recon- isopach map, is consistent with a northeast-oriented
struction of Uliana et al. (1989), which envisages a Basin extensional stress field (Figure 7A).
and Range type process of extension driven by orogenic The Jurassic postrift basin was largely an underfilled
collapse. marine basin into which deltaic and fluvial depositional
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 391

Figure 8—Paleogeographic reconstruction following Callovian inversion. (A) Basal continental and marine sedimentary
cover of the lower Lotena Group. (B) Upper Lotena deposition of evaporites and limestones in a structurally closed basin.

systems prograded. Despite this freshwater influx, It was also a precursor to more massive latest Oxfordian–
marine invertebrate assemblages indicate normal marine early Kimmeridgian inversion along the Huincul arch
salinities (Riccardi, 1983). In the western part of the basin and the start of a new phase of basin subsidence.
(e.g., Vaca Muerta range), Lajas Formation sedimentary
facies suggest high-energy shelf and mesotidal ranges Late Jurassic Inversion Tectonics
(Dean, 1987), implying a paleogeography that was
partially open to the west. The provenance for the deltaic General uplift and erosion affected much of the
and fluvial sediments was toward the east and southeast Neuquén basin during the Callovian, especially the
(Uliana and Legarreta, 1993). Seismic sections record the Dorsal de Huincal area and the western and south-
offlapping progradational geometries, while exploration western parts of the basin (Figure 8A). In the north-
wells demonstrate the lithologic associations. The postrift western parts of the basin, inversion resulted in erosion
sequence comprises the upper Cuyo Group and its of upper Bajocian–lower Callovian strata (Dellapé et al.,
constituent Loma Negra, Lajas, Challaco, and Punta 1979; Gulisano, 1981).
Rosada formations (Figure 2). These grade into or Inversion of rift-related fault blocks was particularly
interfinger with Los Molles topset beds. By the end of conspicuous along the Dorsal de Huincul, but also
Cuyo deposition, the Neuquén basin was increasingly affected other extensional tracts. The continental deposits
silled and had a restricted circulation, as evidenced by in the basal part of the Lotena Formation were derived
the Tábanos evaporites (Figure 4D). Figure 2 shows that from inverted fault blocks and deposited as synorogenic
inversion was possibly initiated in the Callovian, molasse (Figure 8A). Partial erosion of upper Cuyo strata
suggesting a depositional response (evaporites and lime- occurred along the margins of the basin. These rocks are
stones) to inversion-assisted shallowing. absent in the Cordillera del Viento and in the Reyes and
Cara Cura ranges (Figure 4A) (Gulisano, 1981). In the
Dorsal de Huincul, it is difficult to estimate the amount
of Callovian erosion because younger erosional events
LATE JURASSIC–CRETACEOUS were even more substantial.
SUBSIDENCE At the end of this period of inversion, marine to evap-
oritic facies were deposited in a widespread downwarp
An early Callovian unconformity marks a conspic- that we attribute to relaxation of compressional stresses.
uous change in the style of sedimentation to the This initial sedimentary drape is reflected in the upper
evaporite-limestone dominated Lotena Group (Figure 2). Lotena Formation. These deposits shoal upward into
392 Vergani et al.

Figure 9—Late Jurassic Araucanian inversion which focused on the southern extensional tract. (A) Structural inversion and
partitioning by northwest-directed transfer faults. (B) Late Jurassic Tordillo paleogeography.

evaporitic rocks of the Loteniano-Chacayano sequence, inversion, nor was the extensional system to the south of
which includes La Manga, Barda Negra, and Auquilco the inverted belt (El Sauce and Río Negro trend, Figure
formations (Figures 2, 8B). 4A). This selective deformation is characteristic of the
A second and even more intense period of inversion inversion that covers an area 300 km long and 50 km
occurred in the late Oxfordian and earliest Kimmerid- wide (Figure 9A).
gian. This deformation, known as the Araucanian event, The transfer faults acted as structural boundaries
was focused on the old Piedra del Aguila–Barda during inversion. In the Plaza Huincul area, tectonic
Colorada–Chacaico trough and extended eastward to inversion was segmented by northwest-trending transfer
Estancia Vieja (Figures 4A, 9A). The inversion process faults that differed from the southwest-directed
was typically selective, avoiding some extensional struc- extension of the Early Jurassic (Figures 7A, 9A),
tures while reactivating others. Besides inversion of the suggesting that this Late Jurassic inversion was a
Dorsal de Huincul tract and the southern and south- response to fundamental reorganization of regional
western parts of the basin, isolated structures were stress fields. These Late Jurassic transfer faults offset and
uplifted in the Chacaico, Vaca Muerta, Reyes, and Cara separated some areas of inversion from others that were
Cura ranges (Figure 4A) where Cuyo and Lotena succes- unaffected by inversion. An example is the Barda
sions were deeply eroded. Colorada half-graben, which was intensely inverted
Figures 10 through 14 show the character and dimen- relative to the neighboring El Sauce half-graben (Figures
sions of this Oxfordian–Kimmeridgian deformational 10, 12); the latter was not deformed. The change from
event. Over 2000 m of Choiyoi, pre-Cuyo, and Cuyo compression to extension within this area has been inter-
Group sedimentary rocks were partially eroded in the preted as the result of transpression and transtension
Plaza Huincul area (Ploszkiewicz et al., 1984). North- within a wrench fault system (Orchuela et al., 1981;
northwest directed transfer faults divided this inverted Ploszkiewicz et al., 1984). We suggest that an initial rift
belt into three sectors (Figure 9A), each showing phase was followed by repeated tectonic inversions.
decreasing deformation toward the east (compare Structural inversion significantly restricted the
Figures 10, 13, 14). In the west, this phase affected the Neuquén basin. Fluvial deposits at the base of the
Vaca Muerta, Chacaico, and Cara Cura ranges by trunca- Tordillo Formation are thickest in the west where they
tion of the Cuyo and Lotena Group (Figure 11). In are augmented by large volumes of pyroclastic material
Estancia Vieja (Figure 14), the northwest-oriented (Gulisano, 1981). Thinner fluvial units are preserved
inversion only affected the east-west fault trend between uplifted fault blocks in the southwestern part of
bordering the pre-Cuyo and Cuyo depocenters. The the basin. Tectonic quiescence and a low-relief source
Bajada Vidal extensional fault was not affected by the terrain resulted in stratigraphic onlap of the Tordillo
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 393

Figure 10—Uninterpreted and interpreted seismic section showing structural inversion of the western part of Dorsal de
Huincul in the Barda Colorada area. Vertical exaggeration is 2:1 at 1 sec (two-way time). See Figure 3 (line A) for location.

across basin margin ramps and intrabasin highs (Figure Andico cycle (Figure 2) (Gulisano et al., 1984a). Mitchum
9B). In the southeastern parts of the basin, the sediments and Uliana (1985) describe the progradational strati-
are typically finer grained and are attributed to lacustrine graphy of this Mendoza sequence, including the prolific
and eolian paleogeographies (Arregui, 1993); they are Vaca Muerta source rocks (see their figures 5–16).
known as the Catriel and Sierras Blancas formations
(Figure 2) (Marchese, 1971). Early Cretaceous Subsidence
The final phase of continental sedimentation during
the Tithonian coincided with regional subsidence which During the early Valanginian, a major change in the
we attribute to compressional relaxation (see Bally and pattern of basin subsidence resulted in interruption of the
Snelson, 1980). Marine encroachment and expansion of marine sedimentation that characterized the previous
the basin is reflected in the Mendoza Group of the cycle. The western and southern parts of the basin were
394 Vergani et al.

Figure 11—Photograph of strata deformed during the Arau-


canian event, Vaca Muerta range. (Courtesy of F. Drysdale.)

again subjected to inversion and regional uplift. Inversion Figure 12—Interpretation of Dorsal de Huincul inversion of
of older extensional faults was intense in the Dorsal de earlier extensional basins. Inversion is a selective process,
Huincul area and its extension toward the southwest. and in this reconstruction, it has deformed the Barda
Colorada half-graben while the neighboring El Sauce half-
Along the eastern margin of the basin, moderate uplift
graben continued to subside. 1, Paleozoic basement; 2,
resulted in minor stratigraphic truncation (Figures 5, 15). Pre-Cuyo early rift; 3, Lower Cuyo late rift; 4, Upper Cuyo
Seismic data (Figure 13) show truncation of the lower postrift; 5, Lotena Group. Not to scale.
Mendoza Group beneath the Valanginian unconformity
along the northern flank of the Huincul arch. Mendoza
stratigraphy was also eroded along the eastern margin of
the basin (Figures 5, 6). The inverted areas contributed oriented linear trends reflect reactivation of Triassic–
sediments to the interior of the basin where they were Early Jurassic transfer faults in a normal sense (Welsink
preserved to variable thicknesses. Westward, fluvial et al., 1995). These characteristics are consistent with the
sandstones and conglomerates graded laterally into their northwest-directed transfer faults that dissect the Dorsal
distal equivalents in the basin center (Figure 15) (Gulisano de Huincul and mark a reorganization of stress fields.
et al., 1984a). In the east, thin-bedded fluvial deposits of
calcareous sandstones and shales interfingered with
marine shales and carbonates toward the north. This Late Cretaceous Subsidence
clastic succession forms the Mulichinco Formation
(Figure 2). Farther north, carbonates of the Chachao Renewed tectonic activity and inversion during the
Formation were deposited on a platform flanking the earliest Cenomanian was characterized by subtle erosion
passive eastern margin of the basin (Figure 2). and northward thinning of the Rayoso Group in the
Tectonic quiescence characterized this basin from western part of the basin (Figures 2, 16) (Uliana et al.,
Valanginian time onward and resulted in renewed 1975). This event is recognized in outcrop along the
marine transgression and deposition of the platform Dorsal de Huincul and is conspicuous in well and
sediments of the Agrio Formation. These sediments seismic data (Figures 10, 12). Along the eastern margin of
interfingered with prograding fluvial clastics of the the basin, seismic sections show moderate truncation of
Centenario Formation along the margin (Figure 2). A the Rayoso Group and development of a slightly angular
general shallowing of the basin was accompanied by unconformity (Figures 5, 6, 10, 13). Correlation of wells
evaporitic and clastic sedimentation within a realm of and field sections shows that the unconformity is wide-
regional subsidence. This Rayoso Group represents the spread in the northeastern and northwestern parts of the
culmination of the Andico cycle. The Lower Cretaceous basin (Uliana et al., 1975; H. Sosa, 1994, personal commu-
Mendoza-Rayoso succession is punctuated by several nication).
unconformities that define the stacking of multiple trans- This tectonic phase reactivated the provenance areas
gressive-regressive sequences, reflecting intermittent that resulted in accumulation of continental deposits of
subsidence and eustatic processes (Figure 2) (see the Neuquén Group. This sedimentation persisted until
Mitchum and Uliana, 1985; Uliana and Legarreta, 1993). the end of the Cretaceous, at which time the continental
Figure 7B illustrates the Upper Jurassic–Lower Creta- and marine deposits of the Malargüe Group marked the
ceous isopachs that record the style of basin subsidence culmination of the Riográndico cycle (Figure 2)
following post-Araucanian relaxation. Local northeast- (Legarreta and Gulisano, 1989).
Figure 13—Seismic section showing inversion in the central part of Dorsal de Huincul in the Aguada Villanueva area. Vertical
exaggeration is 1.5:1 at 1 sec (two-way time). See Figure 3 (line B) for location.
396 Vergani et al.

Figure 14—Seismic section showing structural inversion in the eastern part of the Dorsal de Huincul in the Estancia Vieja
area. Note deformation decreasing toward the east. Vertical exaggeration is 2.3:1 at 1 sec (two-way time). See Figure 3
(line C) for location.
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 397

Figure 15—Paleogeography of Valanginian Mulichinco Figure 16—Paleogeography of Upper Cretaceous lower


Formation showing inversion-restricted depocenter and Neuquén Group dominated by terrestrial depositional
sediment influx. systems. Cenomanian growth along the Dorsal de Huincul
locally divided the basin.

TERTIARY COMPRESSION AND


SUB-ANDEAN FORELAND BASIN
In the Eocene, the Neuquén basin entered a new
phase of deformation and basin subsidence (Figure 17).
The north-trending Andean fold and thrust belt
encroached on the western margin of the basin,
deforming the Mesozoic prism in front of it (Figure 3).
Preexisting normal faults of the rift phase, as well as the
rock mechanics of the Mesozoic sequence, controlled the
deformation style in this belt. Thick-skinned deformation
involved the Paleozoic–Triassic basement (Figures 18,
19). Deformation is characterized by first-order folds that
were generated by inversion of variably dipping normal
faults (Manceda and Figueroa, 1993). An example is the
structural tract of the Cordillera del Viento and its
southward continuation where the basement and oldest
Jurassic sedimentary rocks outcrop. Structures of the
Reyes and Cara Cura ranges have Triassic basement in
their cores. Various rift faults were reactivated as back-
thrusts during the Tertiary (Figure 19).
South of the Vaca Muerta range, including the
Chacaico and several smaller ranges, basement was reac-
tivated during the various Jurassic–Cretaceous episodes
of inversion. The interference of the thrust belt with the
older inversion zone resulted in a complex pattern of
cross-cutting trends, folds with varying orientations, and
dog-leg types of faults.
The thin-skinned deformation was controlled by the
rock mechanics of the Jurassic–Cretaceous sedimentary Figure 17—Tertiary orogenesis and development of a
cover. Important detachment levels are present within foreland basin.
398 Vergani et al.

Figure 18—Cross section showing fold and thrust belt, frontal syncline, Dorsal de los Chihuidos, and basin margin.
No vertical exaggeration. See Figure 3 (line F) for location.

Figure 19—Cross section showing external fold and thrust belt, frontal syncline, and basin margin. No vertical exaggeration.
See Figure 3 (line G) for location.

the marine shales of the Cuyo and Mendoza groups in overlain by Oligocene basalt flows of variable thickness.
which overpressured zones occur locally. Evaporites of In the southern part of the basin, continental deposits of
the Rayoso Group form a major upper detachment level the Michihuao Formation, possibly of Oligocene–early
in the northern part of the basin. These detachment levels Miocene age, underlie the sandstones and tuffs of the
control the deformation style in the foldbelt and its Collón Cura Formation. Finally, fluvial deposits uncon-
eastern flank (Figures 18, 19). In this important hydro- formably overlie the various Tertiary units. This succes-
carbon province, the interpretation of a triangle zone is sion is known as the Pliocene Tristeza Formation, which
based on seismic and well data (Ploszkiewicz, 1987; is better developed toward the north. Late Tertiary
Viñes, 1990). The thin-skinned structural style developed intrusive events (Huincan and Desfiladero Negro forma-
from north of the basin to the Agrio River in the south. It tions) occurred in both depositories and also affected the
is characterized by en echelon box folds with different Jurassic–Cretaceous stratigraphic column by intrusion of
vergences and local extent. intermediate and basic dikes and sills.
Thrust belt loading produced several internal
depocenters by fault reactivation and flexural subsidence
(Figure 17). In the west, thick volcanic suites known as
the Serie Andesitica are overlain by fluvial and lacustrine DISCUSSION
sedimentary deposits that are intercalated with tuffs and
andesites of late Oligocene–early Miocene age (Uliana, The tectonic evolution of the Neuquén basin involved
1978; Rapela and Kay, 1988). Miocene clastic and pyro- several episodes of extensional subsidence and structural
clastic deposits of the Collón Cura Formation follow inversion as the Patagonian lithosphere adjusted to two
unconformably (Figure 2). A Recent basaltic and major reorganizations of regional stress fields. This basin,
fluvioglacial succession covers much of this area. together with other Mesozoic basins of Patagonia, is
Discontinuous fluvial and intercalated volcaniclastic tectonically linked to Paleozoic precursors. First, many of
sediments are present at the foot of the external fold and these basins overlie Paleozoic depocenters, suggesting
thrust belt. During the Eocene–early Oligocene, clastic that they developed by reactivation of preexisting struc-
and pyroclastic sediments of the Carrere Formation were tures (see Fernandez-Seveso and Tankard, 1995, their
deposited (Uliana, 1978). These are unconformably figure 1). Second, Middle Triassic–Early Jurassic exten-
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 399

Figure 20—(A) Northeast-directed transfer faults accommodated extension of the Triassic–Early Jurassic rift basins. (B) The
change to a northwest-directed σ3 reactivated these transfer faults in a normal sense, as shown by local northeast-oriented
isopach trends. In contrast, faults subparallel to the extension direction were reactivated as transfer faults.

sion is attributed to extensional collapse of the Permian–- al., 1989; Tankard et al., 1995; Welsink et al., 1995). Figure
Triassic orogens (Tankard et al., 1995). Dewey (1988) 4A shows a network of extensional faults that formed
relates extensional orogenic collapse to overthickened Triassic–Early Jurassic rift basins. The northwest trends
crust beneath orogenic belts and structural anisotropy. (e.g., Entre Lomas and Bajada Vidal) and the north-south
Extension of these Triassic–Jurassic basins in southern structural trends (e.g., Catan Lil, Chihuidos, and Reyes)
South America was accomplished by northeast-oriented imply a σ3 oriented toward the southwest. The pattern of
transfer faults (Welsink et al., 1995, their figure 17). These fault offsets, different scales of extensional tracts, variable
faults were parallel to the extensional stress at that time structural styles, and northeast-trending structural
(Figure 20A). A Basin and Range type of extensional boundaries or sidewalls suggest that the Triassic–Early
mechanism (see Wernicke, 1985; Verrall, 1989) is inferred Jurassic period of rifting involved northeast-oriented
for these early rifts. transfer faults. From Aalenian time onward, fault-
The Early Jurassic Neuquén succession contains controlled subsidence gradually decreased and the
diverse assemblages of invertebrate faunas that suggest discrete rifts and their intervening basement highs were
normal marine salinities and circulation. The volumi- yoked together in a broad postrift basin that subsided
nous freshwater influx reflected in the alluvial Challaco uniformly. This Triassic–Jurassic history of rift and
and Punta Rosada formations was insufficient to dilute postrift subsidence describes the first stage of basin
the basin, suggesting a direct westward connection to the subsidence spanning about 50 m.y. and is typical of
Pacific. This interpretation is supported by mesotidal extension.
sedimentation of some Lajas facies (Dean, 1987). These From Callovian time on, the Neuquén basin was
mesotidal ranges, marine affinities of the invertebrate subjected to periodic compressional reactivation of the
faunas, and volcanic materials in the Cuyo succession earlier extensional structures, locally creating massive
together support a back-arc setting for the Neuquén inversion structures. The largest of these, and the most
basin behind an irregular volcanic chain that permitted prominent structural element in the Neuquén basin, is
periodic exhange of water. the Dorsal de Huincul, or Huincul arch. The Late Jurassic
The structural framework of the Neuquén basin is inversion marks a fundamental reorganization of
complex because it has been repeatedly reactivated regional stress fields and the start of a new phase of basin
through different phases of Mesozoic extension and subsidence. The principal diastrophism and inversion
overprinted by several episodes of structural inversion. along the Huincul trend dates to the late Oxfordian–
Regional relationships provide a useful guide (Uliana et earliest Kimmeridgian. Seismic data show that variable
400 Vergani et al.

geometries of inversion along strike on the Huincul tectonics, with the climax of mountain building in the
structure were compartmented by regular northwest- Miocene and Pliocene. The importance of this Andean
trending offsets that suggest transfer faults of that orogeny was its influence on structural trap formation by
orientation. Post-Kimmeridgian deepening of the fault reactivation, especially in the northwestern part of
Neuquén basin was progressive toward the northwest the basin, and the blanketing effect of the thick foreland
(Figure 20B). The coincidence of local northeast- basin cover on source rock maturation.
trending isopachs and Triassic–Jurassic transfer faults There is obvious structural control on the distribution
suggest that this deepening was the result of the reacti- of oil and gas fields (compare Figures 3B, 4A). The giant
vation of transfer faults in a normal sense. Although in Loma La Lata oil and gas field, for example, is controlled
some cases faults cut as high as the Quintuco Formation, by intersecting northeast- and northwest-oriented faults
most of the fault activity is concealed beneath the thick along which an inversion anticline has formed. Else-
blanket of postrift sediments. Broad anticlinal warping where, the structural relationships of inversion and
and flexuring are evidence of deep-seated tectonic erosion are an important control of field distribution.
activity. Most oil and gas fields have a structural component of
This new structural framework precisely matches the trapping. Purely stratigraphic traps are less significant.
style of Late Jurassic–Cretaceous extension that preceded
fragmentation of southwestern Gondwana and the
opening of the South Atlantic. Extension along the
southern margin of South Africa and the eastern Acknowledgments We thank R. Allen, C. Arregui, F.
seaboard of Argentina and Uruguay (e.g., Santa Lucia, Drysdale, D. Lehto, D. Loureiro, R. Manceda, H. Piana, L.
Punta del Este, Salado, Colorado, and Outeniqua basins) Rebori, H. Sosa, M. Uliana, and R. Viñes for helpful discus-
started in about the Kimmeridgian (Tankard et al., 1995). sions. Typing was done by K. Bojarski and drafting by R.
These extensional tracts were linked to a system of west- Colombres. Finally, we thank the management of Perez
northwest diverging wrench faults. Examples include the Companc, especially R. Blocki and H. Marchese, for support
Agulhas-Malvinas fracture zone and the Martin Garcia and permission to publish.
trend, implying σ3 oriented toward the northwest in Late
Jurassic–Cretaceous time (Welsink et al., 1995). In this
sense, the Late Jurassic Dorsal de Huincul is directly
linked to the reorganization of stress fields that initiated REFERENCES CITED
Atlantic extension. The Huincul inversion is a conspic-
uous focal point of this episode of extension. Argentinian Petroleum Institute, 1993, Bolétin estadistico 1992
Extensional fault-controlled subsidence was largely Enero/Abril 1993: Buenos Aires, p. 1–24.
restricted to the Late Triassic–Early Jurassic in the Arregui, C., 1993, Análisis estratigráfico-paleoambiental de la
Neuquén basin. Nevertheless, comparable rates of subsi- Formacion Tordillo en el subsuelo de la cuenca neuquina:
dence were repeated in the Late Jurassic and earliest Décimo Segundo Congreso Geológico Argentino y
Segundo Congreso de Exploración de Hidrocarburos,
Cretaceous when the Tordillo–Vaca Muerta sequence Mendoza, v. 1, p. 165–169.
accumulated. We attribute this to relaxation of in-plane Bally, A. W., and S. Snelson, 1980, Realms of subsidence, in A.
stresses (see Bally and Snelson, 1980). This structural D. Miall, ed., Facts and principles of world petroleum
complexity resulted in at least two prolific source rock occurrence: Canadian Society of Petroleum Geologists
intervals: the Los Molles and Vaca Muerta shales. The Memoir 6, p. 9–94.
early Neuquén basin was structurally compartmented, Carbone, C., 1988, Sismoestratigrafía del Grupo Cuyo inferior
suggesting that Los Molles source rocks may have accu- en la cuenca neuquina: Boletin de Informaciones Petrol-
mulated in a variety of rift depocenters, not all of which eras, Tercera Epoca, v. 16, p. 67–87.
were connected (Figure 4). Geochemical analysis suggests Cloetingh, S., A. J. Tankard, H. J. Welsink, and W. A. M.
marine as well as lacustrine affinities (Zumberge, 1993). Jenkins, 1989, Vail’s coastal onlap curves and their correla-
tion with tectonic events, offshore eastern Canada, in A. J.
This study broadly agrees with the interpretation of Tankard and H. R. Balkwill, eds., Extensional tectonics and
Uliana and Legarreta (1993) which ascribes the Neuquén stratigraphy of the North Atlantic margins: AAPG Memoir
basin to Triassic–Early Jurassic extension followed by a 46, p. 283–293.
long period of Middle Jurassic–Paleogene subsidence. Cobbold, P., A. C. Massabie, and E. A. Rossello, 1986,
Where we differ is in suggesting that regional subsidence Hercynian wrenching and thrusting in the Sierras
was more intermittent than uniform and was punctuated Australes foldbelt, Argentina: Hercynica, v. 2, p. 135–148.
by several episodes of structural inversion. There is a Dean, J. S., 1987, Depositional environments and paleogeog-
broadscale agreement of the Neuquén sequence stratig- raphy of the Lower to Middle Jurassic Cuyo Group,
raphy with the “global” sea level chart of Haq et al. Neuquén basin, Argentina: Ph.D. dissertation, Colorado
(1987) (Figure 2). However, in a dynamic tectonic setting, School of Mines, Golden, Colorada, 586 p.
Dellapé, D. A., C. Mombrú, G. A. Pando, A. C. Riccardi, M. A.
we are unable to separate any eustatic overprint from Uliana, and G. E. G. Westermann, 1979, Edad y correlación
tectonic processes. Suffice it to say, eustasism may de la Formación Tábanos en Chacay Melehue y otras local-
indeed have been a secondary response to tectonism (see idades de Neuquén y Mendoza: Obra Centenario, Museo
Cloetingh et al., 1989). La Plata, v.5, p. 81–105.
The final innovation in basin formation involved Late Dewey, J. F., 1988, Extensional collapse of orogens: Tectonics,
Cretaceous–Tertiary onset of Andean compressional v. 7, p. 1123–1139.
Tectonic Evolution and Paleogeography of the Neuquén Basin, Argentina 401

Digregorio, J. H., and M. A. Uliana, 1980, Cuenca neuquina: Octavo Congreso Geológico Argentino, San Luis, v. 3,
Segundo Simposio de Geologia Regional Argentina, p. 281–293.
Academia Nacional de Ciencias, Córdoba, v. 2, p. 985–1032. Ploszkiewicz, J. V., 1987, Las zonas triangulares de la faja
Fernandez-Seveso, F., and A. J. Tankard, 1995, Tectonics and fallada y plegada de la cuenca neuquina, Argentina:
stratigraphy of the late Paleozoic Paganzo basin, western Décimo Congreso Geológico Argentino, v. 1, p. 177–180.
Argentina, and its regional implications, in A. J. Tankard, R. Ploszkiewicz, J. V., I. A. Orchuela, J. C. Vaillard, and R. F.
Suarez, and H. J. Welsink, Petroleum basins of South Viñes, 1984, Compresión y desplazamiento lateral en la
America: AAPG Memoir 62, this volume. zona de falla Huincul, estructuras asociadas, provincia del
Gulisano, C. A., 1981, El Ciclo Cuyano en el norte de Neuquén Neuquén: Noveno Congreso Geológico Argentino, San
y sur de Mendoza: Octavo Congreso Geológico Argentino, Carlos de Bariloche, v. 2, p. 163–169.
San Luis, v. 3, p. 579–592. Rapela, C. W., and S. M. Kay, 1988, Late Paleozoic to Recent
Gulisano, C. A., and G. A. Pando, 1981, Estratigrafía y facies de magmatic evolution of northern Patagonia: Episodes, v. 11,
los depósitos jurásicos entre Piedra del Aguila y Sañico, p. 175–182.
Departamento Collón Curá, Provincia del Neuquén: Riccardi, A. C., 1983, The Jurassic of Argentina and Chile, in M.
Octavo Congreso Geológico Argentino, San Luis, v. 3, Moullade and A. E. M. Nairn, eds., The Phanerozoic
p. 553–577. geology of the world II, the Mesozoic, B: Amsterdam,
Gulisano, C. A., A. R. Gutierrez Pleimling, and R. E. Digre- Elsevier, p. 201–263.
gorio, 1984a, Análisis estratigráfico del intervalo Titho- Riccardi, A. C., S. E. Damborenea, and G. E. G. Westermann,
niano-Valanginiano (formaciones Vaca Muerta, Quintuco y 1990, Middle Jurassic of South American and Antarctic
Mulichinco) en el suroeste de la Provincia del Neuquén: Peninsula, in G. E. G. Westermann and A. C. Riccardi, eds.,
Noveno Congreso Geológico Argentino, San Carlos de Jurassic taxa ranges and correlation charts for the Circum-
Bariloche, v. 1, p. 221–235. Pacific: Newsletters on Stratigraphy, v. 21, p. 105–128.
Gulisano C. A., A. R. Gutiérrez Pleimling, and R. E. Digre- Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M.
gorio, 1984b, Esquema estratigráfico de la secuencia jurásica Turic, A. B. Franca, E. J. Milani, B. B. de Brito Neves, N.
del oeste de la provincia de Neuquén: Noveno Congreso Eyles, J. Skarmeta, H. Santa Ana et al., 1995, Tectonic
Geológico Argentino, San Carlos de Bariloche, v. 1, controls of basin evolution in southwestern Gondwana, in
p. 236–259. A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of basins of South America: AAPG Memoir 62, this volume.
fluctuating sea levels since the Triassic: Science, v.235, Uliana, M. A., 1978, Estratigrafía del Terciario: Séptimo
p. 1156–1167. Congreso Geológico Argentino, Neuquén, Relatorio,
Hinterwimmer, G. A. and J. M. Jauregui, 1984, Análisis de p. 67–83.
facies de los depósitos de turbiditas de la Formación Los Uliana, M. A, and L. Legarreta, 1993, Hydrocarbons habitat in
Molles en el sondeo Barda Colorada Este, Provincia del a Triassic-to-Cretaceous sub-Andean setting: Neuquén
Neuquén: Noveno Congreso Geológico Argentino, San basin, Argentina: Journal of Petroleum Geology, v. 16,
Carlos de Bariloche, v. 5, p. 124–135. p. 397–420.
Kay, S. M., V. A. Ramos, C. Mpodozis, and P. Sruoga, 1989, Uliana, M. A., D. A. Dellapé, and G. A. Pando, 1975, Distribu-
Late Paleozoic to Jurassic silicic magmatism at the ción y génesis de las sedimentitas Rayosianas (Cretácico
Gondwana margin: analogy to the middle Proterozoic in inferior de las Provincias de Neuquén y Mendoza,
North America: Geology, v. 17, p. 324–328. República Argentina): Segundo Congreso Ibero-Americano
Legarreta, L., and C. A. Gulisano, 1989, Análisis estratigráfico Geologia Económica Actas, v. 1, p. 151–176.
secuencial de la cuenca neuquina (Triásico superior- Uliana, M. A., K. T. Biddle, and J. Cerdán, 1989, Mesozoic
Terciario inferior), Argentina, in G. A. Chebli and L. A. extension and the formation of Argentine sedimentary
Spalleti, eds., Cuencas sedimentarias Argentinas: Serie basins, in A. J. Tankard and H. R. Balkwill, eds., Extensional
Correlación Geológica 6, p. 221–243. tectonics and stratigraphy of the North Atlantic margins:
Legarreta, L., and M. A. Uliana, 1991, Jurassic–Cretaceous AAPG Memoir 46, p. 599–614.
marine oscillations and geometry of back-arc basin fill, Verrall, P., 1989, Speculations on the Mesozoic–Cenozoic
central Argentine Andes, in D. I. M. McDonald, ed., Sedi- tectonic history of the western United States, in A. J.
mentation, tectonics and eustasy: International Association Tankard and H. R. Balkwill, eds., Extensional tectonics and
of Sedimentologists Special Publication 12, p. 429–450. stratigraphy of the North Atlantic margins: AAPG Memoir
Manceda, R., and D. Figueroa, 1993, La inversión del rift 46, p. 615–631.
mesozoico en la faja fallada y plegada de Malargüe, Viñes, R. F., 1990, Productive duplex imbrication at the
Provincia de Mendoza: Décimo Segundo Congreso neuquina basin thrust belts, Argentina, in J. Letouzey, ed.,
Geológico Argentino y Segundo Congreso de Exploración Petroleum and tectonics in mobile belts: Paris, Editión
de Hidrocarburos, v. 3, p. 219–232. Technip, p. 69–80.
Marchese, H. G., 1971, Litoestratigrafia y variaciones faciales Welsink, H. J., E. Martinez, O. Aranibar, and J. Jarandilla, 1995,
de las sedimentitas mesozoicas de la Cuenca Neuquina, Structural inversion of a Cretaceous rift basin, southern
Provincia de Neuquén, República Argentina: Asociación Altiplano, Boliva, in A. J. Tankard, R. Suarez, and H. J.
Geológica Argentina Revista, v. 26, p. 343–410. Welsink, Petroleum basins of South America: AAPG
Mitchum, R. M., and M. A. Uliana, 1985, Seismic stratigraphy Memoir 62, this volume.
of carbonate depositional sequences, Upper Jurassic–Lower Wernicke, B., 1985, Uniform-sense normal simple shear of the
Cretaceous, Neuquén basin, Argentina, in O. R. Berg and D. continental lithosphere: Canadian Journal of Earth Sciences,
G. Woolverton, eds., Seismic stratigraphy II: an integrated v. 22, p. 108–125.
approach to hydrocarbon exploration: AAPG Memoir 39, Zumberge, J. E., 1993, Organic geochemistry of Estancia Vieja
p. 55–274. oils, Río Negro norte block, in M. H. Engel and S. A. Macko,
Orchuela, I., J. V. Ploszkiewicz, and R. Viñes, 1981, Reinter- eds., Organic geochemistry: New York, Plenum Press,
pretación estructural de la denominada “dorsal neuquina”: p. 461–471.
402 Vergani et al.

Authors’ Mailing Addresses


G. D. Vergani
H. J. Belotti
H. J. Welsink
Perez Companc
J.J. Lastra Sud 6000
8300 Neuquén
Argentina

A. J. Tankard
Tankard Enterprises
P.O. Box 81002
Calgary, Alberta T2J 7C9
Canada
Hydrocarbon Accumulation in an Inverted
Segment of the Andean Foreland :
San Bernardo Belt, Central Patagonia

G. O. Peroni G. Laffitte
A. G. Hegedus YPF S.A.
J. Cerdan Buenos Aires, Argentina

L. Legarreta
M. A. Uliana
ASTRA C.A.P.S.A.
Buenos Aires, Argentina

Abstract

T he San Bernardo (“Bernárdides”) structural province is a multiply deformed belt transecting the peri-
Andean segment of the Argentine Patagonia. It is distinctly separate from the fold and thrust belt along
the western South American continental plate. The structured zone encompasses a NNW-SSE trending band
about 600 km long and 100 km wide. The area includes faults and folds that involve Precambrian–middle
Paleozoic basement, upper Paleozoic–Jurassic terrestrial to marine sedimentary and volcanogenic wedges, and
Cretaceous nonmarine fill of the intracratonic San Jorge basin. The Cretaceous cover is dominated by discontin-
uous, narrow, box-shaped folds associated with east- and west-verging reverse faults. Oil finds are restricted to
the low-lying unbreached segment between the Senguerr and Deseado rivers where anticlinal structures
developed by contractional reactivation of preexisting normal and strike-slip faults.
Oil generation is attributed to amorphous, largely algal-derived organic matter (TOC, 1–3 wt. %) formed in
brackish to alkaline stratified lakes. Modeling suggests that oil generation occurred from 110 to 30 Ma. Low-
gravity oil (15–25˚ API) resulted from biodegradation and washing. The reservoir comprises alluvial, channel,
and meander belt facies and multistory sandstone sheets. Stacked pay intervals are separated by shales, which
limit interconnectedness in the fields. Porosity loss is due to authigenic zeolites and devitrified glass by-
products in a volcaniclastic grain framework. Traps were formed in the Miocene by compression and inversion
of Jurassic half-grabens, expressed in local pop-ups, folds, and weakly inverted structures. Local uplift has
resulted in erosional removal and breaching of some traps. Hydrocarbon migration was facilitated by a
normally charged system and vertical drainage during the first phase of migration. The sandstone and tuffa-
ceous shale generally impeded migration. Late inversion processes favored hydrocarbon scattering.

Resumen

L a provincia estructural de San Bernardo (“Bernárdides”) es un cinturón polideformado que atraviesa el


sector peri-andino de la Patagonia argentina. Se presenta físicamente separado de la faja plegada y
corrida que bordea el márgen occidental de la placa Sudamericana. La región estructurada comprende una
banda de orientación NNW-SSE que se extiende con un largo superior a los 600 km y un ancho casi siempre
inferior a los 100 km. El área considerada presenta fallas y pliegues que involucran al basamento Precámbrico-
mesopaleozoico, a cuñas sedimentarias terrígenas y volcaniclásticas del Paleozoico Superior y Jurásico, y al
relleno no-marino de la cuenca intracratónica del Golfo San Jorge. La cobertura cretácica presenta un estilo
estructural dominado por pliegues discontínuos de tipo “cajón”, asociados con fallas inversas vergentes al este
y al oeste. Los descubrimientos de petróleo se encuentran confinados al sector deprimido y protegido de la

Peroni, G. O., A. G. Hegedus, J. Cerdan, L. Legarreta, M. A. Uliana, and G. Laffitte, 1995, 403
Hydrocarbon accumulation in an inverted segment of the Andean foreland: San
Bernardo belt, central Patagonia, in A. J. Tankard, R. Suárez S., and H. J. Welsink,
Petroleum basins of South America: AAPG Memoir 62, p. 403–419.
404 Peroni et al.

erosión entre los rios Senguerr y Deseado, donde las estructuras anticlinales se desarrollaron como conse-
cuencia de la reactivación compresiva de fallas de rumbo y normales preexistentes.
La Generación de Petróleo es atribuída a materia orgánic amorfa, mayormente derivada de algas (TOC
1–3%), formada en lagos estratificados de tipo salobre y alcalino. El modelado geoquímico sugiere que la
generación de petróleo comenzó alrededor de los 110 Ma y continuó hasta los 13 Ma. Petróleos densos (15-25º
API) son el resultado de biodegradación y lavado. El Reservorio comprende facies aluviales, de canales y fajas
meandrosas, y cuerpos arenosos laminares de tipo multiepisódico. Es común la presencia de reservorios
apilados, separados por lutitas que limitan la interconexión de capas dentro de los campos. Las limitaciones en
la porosidad se relacionan con ceolitas autigénicas y otros minerales relacionados a la devitrificación de trizas
volcánicas. Las Trampas consisten en pliegues anticlinales, “pop-ups” localizados y estructuras débilmente
invertidas, que se formaron en el Mioceno como consecuencia de compresión regional e inversión de los hemi-
grábenes jurásicos. Algunos alzamientos localizados provocaron remoción erosiva y desventramiento de
algunas de las trampas. La Migración de los hidrocarburos se relaciona con un sistema de carga normal y con
drenaje vertical durante la primera fase de movilización. Las areniscas de tipo discontínuo y las pelitas tobáceas
dificultaron la migración, y el carácter tardío de la inversión estructural favoreció la redistribución dispersiva de
los hidrocarburos.

INTRODUCTION
This paper describes the geology and hydrocarbon
habitat of the San Bernardo belt, a structured segment of
the central Patagonian foreland that forms the western
edge of the Cretaceous San Jorge Basin (Figure 1) (Lesta,
1968; Lesta and Ferello, 1972; Fitzgerald et al., 1990). The
oil generation-migration system and trapping style
depart markedly from classic sub-Andean patterns in
which hydrocarbons are associated with marine source
rocks and subtle structural traps superimposed on a
mildly sloping foreland ramp. In contrast, the larger
fields in the buried part of the San Bernardo deformed
belt appear to be linked to lacustrine source rocks and to
high relief structures shaped by Neogene compressional
inversion of pre-Cretaceous extensional fault systems.
The study area in the Chubut and Santa Cruz
provinces dominates the meseta landscape of central
Patagonia. This area is known as the Patagonides (Keidel,
1925) or the San Bernardo foldbelt. The San Bernardo is
characterized by a NNW-SSE striking band of compres-
sional structures more than 600 km long and about 100
km wide (Figure 1). From the western margin of the
Somuncura massif, it extends southward, crosses the
western margin of the San Jorge basin, and finally reaches
the central part of the Deseado massif. Its eastward distri-
bution is restricted by the little-deformed crust under-
lying the South American Atlantic margin (Urien and
Zambrano, 1973). Toward the west, the structural belt is
confined and separated from the Andes by a 150–250 km
wide tract of little-deformed foreland. The most
depressed and least eroded part of the Bernardides belt
occurs between the Senguerr and Deseado river valleys
(Figure 2), where commercial oil pools have been found.
From a hydrocarbon perspective, the study area is Figure 1—Regional map of southern Argentina and Chile
adjacent to and has much in common with the most showing the location of the San Bernardo belt, principal
sedimentary basins (gray), and mountain ranges (shown
prolific of the Argentine oil provinces, the Cretaceous San by crystalline rock symbol). Hachures show location of the
Jorge basin. Production in the San Jorge began in 1907. San Bernardo belt. The rectangle outlines the area shown
Since then, over 50 medium to large fields and many in Figure 2.
smaller fields have been developed. Cumulative oil
production is greater than 350 million m3 (2.2 billion bbl).
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 405

STRATIGRAPHIC FRAMEWORK
Economic Basement
Igneous and sedimentary rocks in central Patagonia of
pre-Middle Triassic age are generally assigned to the
“basement” by the oil industry. These rock assemblages
include a variety of deformed metamorphic units
reflecting early–middle Paleozoic sedimentation at the
Pacific margin of Gondwana. They have been affected by
the subsequent orogenic evolution of the Patagonia
region (Miller, 1976; Hervé et al., 1981; Hervé, 1988;
Hervé and Mpodozis, 1990; Gonzalez Bonorino, 1991),
late Paleozoic intrusive events (Lesta et al., 1980) with
magmatic arc affinities (Forsythe, 1982), and Carbonif-
erous–Permian marine and nonmarine deposits attrib-
uted to fore-arc and back-arc settings (Ramos, 1983; Gust
et al., 1985) (Figure 3).

Lower–Middle Jurassic
A suite of terrigenous clastic and volcanic rocks as
thick as 2500 m is confined to north-northwest oriented
hanging wall extensional troughs, referred to as Lias and
Tobífera or Lonco Trapial (Ugarte, 1966; Lesta and Ferello,
1972). These sedimentary wedges rest on deformed
basement or on the upper Paleozoic succession. The
depocenters are commonly located adjacent to fold-
related thrusts. Substantial facies and thickness changes
are observed in outcrop (Cortiñas, 1984) near bounding
faults. The seismic geometries of these unconformity-
bounded tilted block wedges (Fitzgerald et al., 1990)
suggest that they postdate the onset of extension. The
facies vary widely between subaerial and normal marine
conditions (Musacchio, 1981; Cortiñas, 1984). Several
stratigraphic intervals, especially the upper members,
include volcaniclastic components (e.g., De Giusto et al.,
1980; Cortés, 1990) that were deposited as lahars and
alluvial fans constructed by debris flows; these were fed
into lacustrine systems and restricted marine seaways.
Coarse detritus from basement rocks were shed across
exposed fault scarps.

Upper Jurassic–Neocomian
The 2000–3000-m-thick Upper Jurassic–Lower Creta-
ceous Las Heras Group records the transition from
restricted sedimentation within multiple depocenters to
amalgamation into a more widespread interior basin
system (Fitzgerald et al., 1990). Persistent subaqueous
depositional environments and limited clastic supply are
reflected in the transgressive internal geometry
(Fitzgerald et al., 1990), the predominantly low
sandstone to shale ratio, and the pervasive black
mudstone and carbonate facies (Lesta and Ferello, 1972;
Barcat et al., 1989). A Classopollis-dominated pollen flora,
ostracod faunas, and widespread zeolites indicate the
Figure 2—Schematic map showing main structural existence of large and persistent saline to alkaline lakes
features and oil fields in the central San Bernardo belt. See and a semiarid climate (Van Niewenhuise and Ormiston,
Figure 1 for location. 1989). Occasional foraminifera in some wells (Laffitte and
Villar, 1982) record an episodic connection with the
Pacific realm.
406 Peroni et al.

Figure 3—Chronostratigraphic summary of San Jorge basin. Numbers on the left are ages (Ma) attributed to the principal
sequence boundaries (SB). (Modified after Fitzgerald et al., 1990.)

Middle–Upper Cretaceous occasional sandstone interbeds (Simpson, 1940; Feruglio,


1948, 1949, 1950). A stacked succession of alternating
The Upper Cretaceous Chubut Group consists of a loess (Spalletti and Mazzoni, 1979) and tongues of
500–2000-m-thick succession of monotonously alternating shallow marine deposits suggests a low-relief setting that
mudstones and sandstones with high pyroclastic influx was episodically flooded by Atlantic waters during
(Teruggi and Rosetto, 1963). Lateral facies and thickness periods of eustatic highstand (Legarreta et al., 1990).
changes are gradual (Brown et al., 1982; Fitzgerald et al., Although previous authors have postulated a Cretaceous
1990). Proximal members contain numerous paleosols, folding event (Keidel, 1925) and multiple episodes of
while their more distal equivalents incorporate mudstone Andean compression (e.g., Feruglio, 1929), the persis-
and cross-bedded sandstone lenses and sheets (Roll, 1938; tently parallel seismic reflection patterns argue against
Sciutto, 1981). These depositional systems are attributed to the occurrence of separate deformation phases. North of
anastomosing and meandering river channels, broad Las Heras (Figure 2), the lower–middle Miocene strata of
alluvial plains, and shallow ephemeral lakes and flood- the Superpatagoniano succession are involved in the
plain swamps (Teruggi and Rosetto, 1963; Sciutto, 1981). Bernardides compressional folding, suggesting that
Open marine influence is only encountered toward the contractional deformation of the foreland began no
top of the succession in the Salamanqueano beds (Feruglio, earlier than the end of the middle Miocene (10–15 Ma).
1948, 1949, 1950). In the Colhue Huapi–Musters lake
district (Figure 2), the Cretaceous tuffaceous sandstones
and mudstones provide a coherent beam that produced
broad buckles and box folds. STRUCTURAL STYLES
The San Bernardo mobile belt displays considerable
Cenozoic along-strike variation in tectonic style, a pattern that
resulted from vertical changes in the nature of prevalent
The Tertiary record is represented by a 500–700-m- deformation and variable levels of erosional exposure.
thick column of pyroclastic-bearing clays and tuffs with The central part of the belt, from central Chubut to north-
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 407

Figure 4—(a) Regional east-west seismic section I and (b) interpreted geologic cross section, showing location of the San
Bernardo belt isolated within a noninverted segment of the Patagonian slab. See Figures 1 and 2 for location.

ernmost Santa Cruz, is conspicuous because of its broad configuration of the fold tract shows limited along-strike
structural features that involve the Cretaceous–lower persistence of the main structural elements. The overall
Cenozoic sequences. The areas north and south of this structure forms a series of blocks characterized by linear
middle segment that span the central Deseado and elements with slightly different prevailing orientations.
western Somoncura massifs expose deeper horizons and Smaller folds and faults with oblique orientations
show the structural attitude at the level of the earlier provide a link between these main segments.
Mesozoic and Paleozoic sedimentary and volcanic Many of the anticlines are sinuous to subparallel. In
successions. The following discussion focuses on the detail, they are discontinuous and irregularly spaced,
central part of the Bernardides where long and arcuate and they locally relate to one another in en echelon or
anticlines, broad synclines, and a variety of associated relay patterns (Feruglio, 1929). Their cross sections show
faults provide opportunities for hydrocarbon exploration relatively short wavelengths and medium to high ampli-
(Figure 4). tudes. Large folds are usually narrow to box shaped with
The best exposures of the fold trend occur in the intact cores, and in most cases, at least one of their flanks
Musters–Colhue Huapi lake region in central Chubut is cut by reverse faults. The surface anticlines have asym-
(Figure 2), where the Bernardides form a distinct group metric profiles, but they lack a consistent vergence
of anticlinal mountains separated by broad synclinal direction. Backward and forward vergences alternate
valleys. South of the Senguerr River “elbow,” the folds along and across the Bernardides belt (Feruglio, 1929;
plunge beneath the level of the Patagonian meseta Sciutto, 1981), and several of the larger anticlines display
(Figure 5), but industrial seismic imaging demonstrates along-strike changes in facing direction.
their continuity as far south as the Deseado River valley. Seismic control demonstrates that steep to slightly
South of that latitude, the northward-oriented folds crop overturned limbs generally evolved into longitudinal
out again over the northern fringe of the Deseado massif. thrusts. Changes in vergence appear to be associated
Along the western side of the belt, transition to the with relay transfers between thrusts located along each
nonfolded adjacent domain is abrupt and defined by a of the flanks. In consequence, sections across fold culmi-
sharp increase in the structural relief (Fitzgerald et al., nations show the core zones as bivergent thrust slices or
1990) along steep fold limbs or reverse faults. The eastern downward-pointing “pop-up” wedges (Figures 6, 7).
margin of the folded zone displays a more gradual loss The southern plunge of the Castillo anticline illustrates
of structural relief across a zone where the folds interfere the nature of along-strike changes in shortening style
with east-west oriented normal fault blocks. The internal (Figure 8). A series of right-lateral tear faults accommo-
408 Peroni et al.

Figure 5—Aerial photograph


showing outcrop pattern and
the southern plunge of the
Codo del Senguerr anticline.
See Figure 2 for location.

date the transfer from an asymmetric buckle into a


westward-dipping thrust fault. Galeazzi (1989) has
reported about the same amount of shortening north and
south of the transfer zone.
Previous studies of the area (Barcat et al., 1984;
Galeazzi, 1989; Meconi, 1989) found that the large anti-
clines display internal complexities related to strike-slip
faults. The long and arcuate Chenque-Challao anticline
(Figure 9) (Barcat et al., 1989), for example, shows consid-
erable internal segmentation due to differential displace-
ment along variably oriented tear faults that we interpret
as Riedel shears. These lower hierarchy fault surfaces
have not been detected in seismic sections, presumably
because of their limited extent and small vertical offset. It
is not known whether these tear zones involve basement
rocks. Working at a regional scale, Barcat et al. (1984)
noticed the existence of en echelon fold patterns and
deduced the presence of northwest-oriented strike-slip
faults trending at an angle to the Bernardides master
alignment. These left-lateral oblique shears do not
obviously cut the Cretaceous succession but are recog-
nized as a result of basement activity along preexisting
faults.

Figure 6—(a) Seismic line II and (b) interpreted geologic


cross section, showing inverted half-graben near Alto Río
Senguerr–Barranca Yankowsky. Numbers in (b) refer to
sequence boundary ages (Ma). See Figure 2 for location.
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 409

Figure 7—Geologic cross section based on seismic line IV showing structural styles that form the anticlines exposed near
Rio Deseado–Cerro Bayo. See Figure 2 for location.

Several authors have attempted to define the amount


and mode of basement participation in the Bernardides
folds (Ferello and Scocco, 1952, in Lesta and Ferello, 1972;
Barcat et al., 1984; Fitzgerald et al., 1990). The existence of
considerable basement relief is demonstrated by drilling
in the region (Ferello and Lesta, 1973), but a more
complete answer will have to await studies of structural
styles in the Deseado and north-central Chubut where
roots of the system come to the surface.
The reflection seismic and well data coverage between
the Senguerr and Deseado valleys provides some clues
to the juxtaposition of contrasting deep and shallow
structures at the level of the basement and the Creta-
ceous cover (Figures 4, 6, 7). Considerable evidence
suggests that the preexisting fault pattern, shaped by
Mesozoic tensional vectors, strongly influenced the
present Bernardides compressional fabrics. Clastic
Figure 8—Schematic map of Castillo anticline. See Figure 2
wedges that formed after Jurassic basin inception
for location. (Simplified after Galeazzi, 1989.)
became contracted and were subjected to incipient to
moderate amounts of tectonic inversion. Most of the
reactivated normal faults show reverse offsets only at and transfer zones) produce a mosaic of blocks and slices
intermediate levels and retain a dominant normal sepa- having slightly different structural attitudes (Cerdán et
ration at depth. Thus, in spite of the impressive upper al., 1990).
level folds and buckles, the amount of total shortening is
thought to be modest.
In striking contrast, the dominant structural style in GEODYNAMIC INTERPRETATION
the noncontracted eastern San Jorge basin is that of a
normal fault block assemblage. The basic architecture The structural geology and development of the oil
consists of 15–25-km-wide structural panels bounded by systems in central Patagonia are best understood in
east-west oriented basement-involved and listric normal terms of the tectonic events that occurred in the final
faults. The activity of these basement fault blocks on a stages of the late Paleozoic Pangea, its Mesozoic breakup,
regional scale is believed to have controlled the deforma- the Cretaceous drift of South America away from Africa,
tion of the Cretaceous–Cenozoic succession. The rupture and finally, the change in boundary conditions that
pattern of the sedimentary cover is one of open to dense induced Cenozoic Andean orogenesis.
upward-branching planar faults rooted in a single The earliest evidence of structural activity along the
basement-penetrating fold (Figure 10). The areal configu- San Bernardo trend was related to the depocenters that
ration of the smaller accommodation faulting is complex, controlled accumulation of the thick upper Paleozoic
particularly at shallow levels. Frequent fault trace discon- successions (Ugarte, 1966; Lesta et al., 1980). Evolution of
tinuities (presumably related to relays, dog-leg offsets, these basins was linked to the tectonic history along the
410 Peroni et al.

been proposed for the proximal side of the Paleozoic


foreland of north-central Argentina. Fernández-Seveso
et al. (1993) and Fernández-Seveso and Tankard (1995)
envision several early Paganzo subsiding troughs as a
distant response to oblique subduction along the South
American margin.
Control of early Mesozoic subsidence (Lesta and
Ferello, 1972; De Giusto et al., 1980) was dominantly
extensional (Table 1) and related to volcanic activity and
emplacement of granitoid stocks (Gust et al., 1985). This
Triassic–Jurassic history fits a framework of supracrustal
gravity collapse (e.g., Dewey, 1988) at the core of a
thermally weakened Paleozoic orogenic welt. Extension
and related volcanic activity became widespread during
the Middle–Late Jurassic, influencing areas well beyond
the San Bernardo belt (Uliana et al., 1985, 1989). Most
authors admit that the Mesozoic Tobífera (Chon Aike,
Lonco Trapial) episode heralded the Africa–South
America decoupling, while crustal extension developed
through a Basin and Range style multicomponent rift
system. Variations in the amount of regional stretching
were accommodated by intracontinental shear zones
aligned with future traces of the mid-Atlantic transform
faults (Windhausen, 1924; Uliana et al., 1989; Rapela,
1990).
By earliest Cretaceous time, the southern South
Atlantic Ocean began to open and fault-driven subsi-
dence started to die out in many Argentine basins
(Uliana et al., 1989). During the Cretaceous and
Paleogene, while South America drifted westward,
intraplate deposits accumulated in a broadening interior
sag. Detailed section balancing (Cerdán et al., 1990)
reveals that gentle extensional faulting, linked to
basement fault systems, segmented the Cretaceous sedi-
mentary cover. As a consequence, the reservoir-prone
Chubutiano suite was dissected by an accommodation
fault network that provided a crucial migration and
trapping element.
Late in the Cenozoic, the extensional stress field was
replaced by compressional conditions leading to
Bernardides folding and inversion, at which time the
structural belt began to emerge as a positive physio-
graphic element. Roughly coeval development and uplift
of the North Patagonian Cordillera (Skarmeta, 1976)
suggests that inversion of the San Bernardo was
promoted by a horizontal stress field generated by differ-
ential interaction at the Nazca–South America plate
boundary. Considering the distance between the
inverted belt and the subduction zone, the notion of
orogenic float (Ziegler, 1987; Oldow et al., 1990; Shaw et
Figure 9—Schematic map of Chenque-Challao anticline. al., 1991) provides a possible mechanism to explain the
See Figure 2 for location. (Modified after Barcat et al., transfer of motion into a spatially segregated transpres-
1984.) sional system that was active behind the late Cenozoic
magmatic arc and the Cordilleran thrust front.
The lack of a well-defined fold vergence suggests that
Pacific margin of Patagonia (Forsythe, 1982; Hervé, 1988; the San Bernardo belt was not deformed as a result of a
Gonzalez Bonorino, 1991). The presence of transtensional unidirectional deformational wave (“bulldozer” mode).
troughs developed along NNW-SSE oriented mega- Instead, the structural styles and stratigraphic relation-
shears might explain the thick Carboniferous and ships within the Tertiary series point toward deforma-
Permian sequences that are confined to linear and tion by regional collapse (“accordion” mode). This
restricted depositional sites. A similar interpretation has involved more or less simultaneous reactivation of older
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 411
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 413

Table 1—Summary of Structural and Depositional Evolution Controlling the Oil System in the San Bernardo Belt

basement faults distributed throughout the contractional observations and wide dispersal of points on Van
zone. The displacement of the focus and the magnitude Krevelen plots reveal a mixture of terrestrial and aquatic
of contractional structuring across the Bernardides were vegetable remains and the presence of type I, II, and III
probably accommodated by strike-slip displacement kerogens, similar to those found in nonmarine basins in
along former graben-bounding and transfer faults. China (Talbot, 1988).
Observations under transmitted light microscopy,
however, demonstrate the dominance of amorphous
organic matter. Organic matter counts show that the
HYDROCARBON HABITAT presence of Celiphus rallus-like nonmarine algae (Figure
11) coincides with the prevalence of amorphous matter
Source Rocks (Figure 12). Where the algae are scarce or absent, the
samples record an increased proportion of terrestrial
Several authors have recognized the hydrocarbon- matter (woody and coaly). The association of algae and
generating potential of the black shale and mudstone in low terrigenous levels suggests that the presence of high-
the Aguada Bandera, Guadal, D 129, and equivalent yield organic matter resulted from peaks in lake produc-
intervals (Laffitte and Villar, 1982; Rodrigo Gainza et al., tivity, attributed to highstand conditions. Several oils in
1984; Yllañez et al., 1989; Van Niewenhuise and the San Jorge basin have been correlated with source
Ormiston, 1989). These Upper Jurassic–Neocomian rocks dominated by amorphous matter and identified as
deposits preserve total organic carbon (TOC) and soluble type II or III kerogens (Yllañez et al., 1989).
organic matter (SOM) levels well above the source On the basis of the deep to shallow water interpreta-
threshold. TOC content ranges between 1 and 2% and tion of the San Jorge lacustrine depositional systems
locally above 3%; SOM is locally above 1000 ppm, and (Fitzgerald et al., 1990) and the apparent vertical changes
total thicknesses may exceed several hundred meters in in organic matter type and richness (G. Laffitte, personal
places. Thickest developments are known at Cerro communication), we suggest a secular change in the
Guadal, Aguada Bandera, Meseta Senguerr, Paso Río productivity and preservation regime. Uppermost
Mayo, Mata Magallanes, and Aguada del León. Visual Jurassic–lower Neocomian organic-rich accumulations
414 Peroni et al.

Figure 11—Photomicrographs of oil-prone organic matter.


(a) (a) Prevalent amorphous organic matter. (b) Elongate form
is a Celiphus rallus-like freshwater algae. (c) Close-up of
the freshwater algae.

(b) (c)

appear to have developed following deeper anoxic and


mildly brackish conditions and in humid climates with
little seasonal contrast. Less prolific upper Neocomian
organic facies appear to represent shallower lakes with
ooidal benches, containing brackish to saline (alkaline)
waters promoted by semiarid climates.

Reservoirs
Hydrocarbon-bearing zones consist of tuffaceous
sandstone strata in the Aptian–Campanian Castillo and
Bajo Barreal formations (Sciutto, 1981). The reservoir
interval thickness spans 1000–2000 m in which the indi-
vidual productive members typically show limited
lateral continuity.
Studies of outcrops adjacent to some of the fields
(Sciutto, 1981; Galeazzi, 1989; Meconi, 1989; Figari et al.,
1990) demonstrate that most of the reservoirs consist of
sandstone and conglomeratic sandstone filling fluvial
paleochannels of various types. Distinctive sandstone
facies include sheet-type accumulations above mobile
channel belts, ribbon deposits linked to fixed channels,
and nonchannelized depositional lobes attributed to
crevasse splaying. They mostly represent multistory Figure 12—Geochemical profile of the organic-rich interval
complexes showing a high degree of internal hetero- in well A. See Figure 2 for location.
geneity. Paleocurrent measurements by Meconi (1989) of
outcrops at the Senguerr River elbow indicate a south- In most fields, the average pay thickness varies
eastward flow direction that shifts to an eastward flow between 7 and 12 m and is usually distributed in three or
toward the younger part of the sequences. These changes four zones. Reservoir quality is mediocre and generally
in flow direction reflect variations in discharge. Earlier inferior to most San Jorge basin reservoirs, a hydro-
conditions of braided patterns and bed load dominance carbon province not noted for performance of its indi-
were replaced by facies attributed to high-sinousity vidual producing zones (Eussler, 1970). Petrographic
rivers and mixed load sedimentation (Galeazzi, 1989). studies by Teruggi and Rosetto (1963) recognize a
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 415

Figure 13—Temperature–burial modeling based on data


from well B. See Figure 2 for location.

mineral assemblage derived from andecite-dacite


materials and the frequent presence of glass shards,
suggesting deposition modified by volcanic activity.
These authors emphasize the presence of apparently
contrasting petrographic characteristics. These
encompass a mixture of angular quartz and plagioclase
clasts, with rounded rock fragments and interstitial
material composed of fragments of different origins and
sizes, that together resemble a graywacke texture. These
textural patterns are overprinted by interstitial analcime
and glass shard montmorillonitization, resulting in low-
performance reservoirs susceptible to formation damage.

Maturation and Migration


Figure 13 shows an attempt to model the time–
temperature evolution of a well located near the center of
the inversion tract. These calculations are in broad
agreement with assessments of present thermal condi-
tions based on vitrinite reflectance measurements. The
results were used to interpret the thermal and migration
histories in the region. The study implies that the
kerogen-rich interval, the oldest of the units with oil-
generating potential, remained in the oil generative zone
Figure 14—Structural geometry contour map showing oil-
for about 80 Ma (Albian–Oligocene). An early entrance bearing anticlinal cluster near Cerro Guadal–Loma del Cuy.
into the oil window (Table 1) implies that, like other See Figure 2 for location.
parts of the San Jorge basin, the oil had ample opportu-
nity to migrate vertically along faults and to become
hosted in pre-Miocene extension-related traps. The (Figure 14). Closures tend to be best defined at shallow
modeling also suggests that late deformation forming the levels and are less common down section. Seismic
present closures postdates the main generating phase control shows that at greater depths the superficial
and may even have formed after the oil had invaded the folding is replaced by faulted structures (Figures 4, 6, 7).
porous Chubutiano strata. Consequently, we speculate Several fields have faulted flanks that have been
that the main effect of structural inversion processes was subjected to variable amounts of structural inversion,
to promote reaccommodation and cause possible suggesting the presence of a whole family of prospects
substantial loss of previously trapped hydrocarbons. that are largely underexplored.
Hydrocarbon distribution patterns within the closures
Hydrocarbon Traps are irregular. Closure sizes are often larger than the oil-
bearing outlines, and hydrocarbon distribution appears
Most of the commercial hydrocarbon occurrences are to be discontinuous within them. These anomalies are
controlled by elongate anticlinal closures that follow an attributed to stratigraphic heterogeneity and local leak
irregular to locally clustered areal distribution of fields zones related to faults having throws below seismic reso-
416 Peroni et al.

lution. Other nonproductive closed structures, such as


the Aguada Bandera anticline, might reflect unfavorable
locations relative to forced relocation of hydrocarbons at
the time of Miocene structural events. As a result of
structural inversion and regional uplift, many folds
located near the northern and southern ends of the
productive tract were eroded down to levels below the
Chubutiano, thus losing their trapping effectiveness.
Although precise figures for cumulative production
and reserve size are not available, we estimate a recovery
level (primary) near 25–30 million m3 for the established
fields. On the basis of these estimates and of the density
of traps expected in inverted belts and an assessment of
the present levels of exploration and delineation in the
region, we suggest an undiscovered potential in the
15–30 million m3 range (Figure 2).

Hydrocarbon Occurrence Figure 15—Cross section based on seismic line III,


showing structural relief and hydrocarbon stratigraphic
Commercial oil fields in the San Bernardo belt occur at distribution at Los Monos–Huetel. See Figure 2 for
depths of 400–1300 m (Figure 15). Pools consist of location.
“black” oil, with gravities between 0.840 and 0.880 g/cm3
and low sulfur content. Gas to oil ratios are usually low,
near 200–700 m3/m3, but some fields have associated structural style is characterized by high-relief structures
free gas accumulations as well. Volume factors are close resembling the deformation patterns typical of some
to 1, and bubble points are lower than the original hydro- compressional belts and intraplate zones of inversion.
static pressures. Reservoirs with significant hydraulic Similarities with the Syrian Palmyrides (Lovelock, 1984;
drive are unknown, and most wells are produced by Best et al., 1990) and the northern African Atlas-Rif
pumping units. Under these circumstances, recovery Mountains (Mattauer et al., 1977; Wildi, 1983) are
factors are low, from 8 to 18%. Common field sizes vary apparent. The Sumatra back-arc system is a prolific oil
from 1 to 2.5 million m3; individual well cumulative province developed in a comparable tectonic setting on
production varies irregularly from 8000 to 25,000 m3. the foreland side of a late Cenozoic orogenic belt that is
linked to a noncollisional convergent plate junction
(Hamilton, 1979; Barber, 1985). Positive structures such
as the “Sunda” type folds of Eubank and Makki (1981)
CONCLUSIONS and Letouzey (1990) form traps related to transpressional
inversion. These have many similarities with our central
The geologic nature and petroleum habitat of the San
Patagonian interpretation.
Bernardo inversion belt are different from the typical
Exploration activity in Chubut and Santa Cruz
sub-Andean petroleum province. Prolific segments of the
suggests that, despite favorable factors such as genera-
South American foreland oil and gas trend, such as the
tion potential and structural traps with substantial
Colombian Los Llanos basin, the Eastern Venezuela
retention capacity, it is unlikely that the Bernardides will
basin (Demaison and Huizinga, 1991), and the
contain fields larger than those already discovered. The
Argentine–Chilean Austral (Magallanes) basin (Pittion
absence of large oil pools is possibly explained by the
and Goudain, 1991), are characterized by laterally
lack of adequate time between structural inversion and
drained petroleum systems that reached the generative
the principal episode of migration (Table 1). Available
stage relatively late during the Cenozoic. Oil and gas
evidence suggests that late inversion resulted in disper-
fields in these basins are located near the base of the
sive reaccommodation of hydrocarbons and partial
largely marine Cretaceous–Tertiary wedge that parallels
degradation of petroleum that was previously trapped
the Andean orogenic belt, in which hydrocarbons are
by extensional faulting such as the large fault block oil
trapped by low-relief structures located updip from the
fields in the prolific San Jorge play belt.
thermally mature source rocks. In contrast, the Bernard-
ides belt is characterized by a predominantly Mesozoic
succession in which both the organic-rich and reservoir
units were deposited in nonmarine settings. Acknowledgments The authors wish to thank ASTRA
From a source rock and reservoir rock perspective, the C.A.P.S.A Exploration and Production for encouragement and
Patagonian succession resembles the depositional styles for permission to publish this paper. Editorial and technical
developed at intraplate hydrocarbon provinces, such as comments by Tony Tankard and an anonymous reviewer are
the eastern China basins (Zaiyi, 1990; Fajing and Shulin, greatly appreciated. Our appreciation is extended to Raúl
1991) and some African basins such as the Sudan (Schull, Genovesi for drafting preparation and to Maria Cristina
1988). The Bernardides are very different in that their Bardelli for typing the manuscript.
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 417

REFERENCES CITED Fernández-Seveso, F., and A. J. Tankard, 1995, Tectonics and


stratigraphy of the late Paleozoic Paganzo basin, western
Argentina, and its regional implications, in A. J. Tankard,
Barber, A. J., 1985, The relationship between the tectonic R. Suarez, and H. J. Welsink, Petroleum basins of South
evolution of southeast Asia and hydrocarbon occurrences, America: AAPG Memoir 62, this volume.
in D. G. Howelr, ed., Tectonostratigraphic terranes of the Fernández-Seveso, F., M. A. Perez, I. E. Brisson, and L. A.
Circum-Pacific region: Circum-Pacific Council for Energy Alvarez, 1993, Sequence stratigraphy and tectonic analysis
and Mineral Resources, Earth Science Series, no. 1, of the Paganzo basin, western Argentina: Comptes
p. 523–528. Rendus XII, J.C.C.-P., Buenos Aires, v. II, p. 223–260.
Barcat, C., J. S. Cortiñas, V. A. Nevistic, N. H. Stach, and H. E. Feruglio, E., 1929, Apuntes sobre la constitución geológica de
Zucchi, 1984, Geología de la región comprendida entre los la región del Golfo San Jorge: Boletín de Informaciones
lagos Musters-Colhue Huapi y la Sierra Cuadrada, Depar- Petroleras, Buenos Aires, no. 63, p. 925–1025.
tamento Sarmiento y Paso de los Indios, Provincia del Feruglio, E., 1948, Descripción geológica de la Patagonia:
Chubut: Noveno Congreso Geológico Argentino, Actas 2, Dirección General de Yacimientos Petrolíferos Fiscales,
p. 263–282. Buenos Aires, Coni, v. 1, 334 p.
Barcat, C., J. S. Cortiñas, V. A. Nevestic, and H. E. Zucchi, Feruglio, E., 1949, Descripción geológica de la Patagonia:
1989, Cuenca Golfo San Jorge, in G. A. Chebli and L. A. Dirección General de Yacimientos Petrolíferos Fiscales,
Spalletti, eds., Serie Correlaciones Geológicas 6: Décimo Buenos Aires, Coni, v. 2, 349 p.
Congreso Geológico Argentino, p. 319–345. Feruglio, E., 1950, Descripción geológica de la Patagonia:
Best, J. A., M. Barazangui, D. Al-Saad, T. Sawaf, and A. Dirección General de Yacimientos Petrolíferos Fiscales,
Gebran, 1990, Bouguer gravity trends and crustal Buenos Aires, Coni, v. 3, 431 p.
structure of the Palmyride Mountain belt and Figari, E. G., J. J. Hechem, and J. Homovc, 1990, Arquitectura
surrounding northern Arabian platform in Syria: Geology, depositacional de las “Areniscas Verdes” de la Formación
v. 18, p. 1235–1239. Bajo Barreal, Provincia del Chubut, Argentina: Tercera
Brown, L. F. ; C. Barcat; W. L. Fisher, and A. Nevistic, 1982, Reunión Argentina de Sedimentología, Actas, p. 130–138.
Seismic stratigraphic and depositional system analysis: Fitzgerald, M. G., R. M. Mitchum, M. A. Uliana, and K. T.
new exploration approaches apply to the Gulf of San Jorge Biddle, 1990, Evolution of the San Jorge basin, Argentina:
basin, Argentina: Primer Congreso Nacional de Hidrocar- AAPG Bulletin, v. 74, p. 879–920.
buros, Petróleo y Gas, Conferencias, Instituto Argentino Forsythe, R., 1982, The late Paleozoic to early Mesozoic
del Petróleo, Buenos Aires, p. 127–156. evolution of southern South America: a plate tectonic
Cerdán, J. J., J. S. Galeazzi, J. D. Enrique, A. I. Jouglard, C. J. interpretation: Journal of the Geological Society of
Perrot, and M. A. Uliana, 1990, Sísmica de detalle mas London, v. 139, p. 671–682.
interpretación integrada: claves para el hallazgo de Galeazzi, J. S., 1989, Análisis de facies y paleocorrientes de la
nuevas reservas en áreas maduras: Tercer Congreso Formación Matasiete en la sección sur del cañadon
Andino de la Industria del Petróleo, Memorias, Tomo I, homónimo: Sierra de San Bernardo, Chubut, Argentina:
p. 27–67. Master’s thesis, Universidad de Buenos Aires, Buenos
Cortés, J. M., 1990, Estratigrafía de las sucesiones volcano- Aires, Argentina, 182 p.
sedimentarias jurásicas del Chubut central, entre Paso de Gonzalez Bonorino, G., 1991, Late Paleozoic orogeny in the
Indios y El Sombrero: Asociación Geológica Argentina, northwestern Gondwana continental margin, western
Revista XLV, v. 1–2, p. 69–84. Argentina and Chile: Journal of South America Earth
Cortiñas, J. S., 1984, Estratigrafía y facies del Jurásico entre Sciences, v. 4, p. 131–144.
Nueva Lubeka, Ferraroti y Cerro Colorado. Su relacióon Gust, D. A., K. T. Biddle, D. W. Phelps, and M. A. Uliana,
con los depósitos coetáneos del Chubut central: Noveno 1985, Associated Middle to Late Jurassic volcanism and
Congreso Geológico Argentino, Buenos Aires, Actas 2, extension in southern South America: Tectonophysics, v.
p. 283–299. 116, p. 223–253.
De Giusto, J. M., C. A. Di Persia, and E. Pezzi, 1980, Hamilton, W., 1979, Tectonics of the Indonesian region:
Nesocraton del Deseado, in J. C. M. Turner, coord., USGS Professional Paper 1078, p. 1–345.
Geología Regional Argentina: Academia Nacional de Hervé, F., 1988, Late Paleozoic subduction and accretion in
Ciencias, Córdoba, p. 1389–1430. southern Chile: Episodes, v. 11, p. 183–188.
Demaison, G., and B. J. Huizinga, 1991, Genetic classification Hervé, F., J. Davidson, E. Godoy, C. Mpodozis, and V.
of petroleum systems: AAPG Bulletin, v. 75, p. 1626–1643. Covacevich, 1981, The late Paleozoic in Chile: stratig-
Dewey, J. F., 1988, Extensional collapse of orogens: Tectonics, raphy, structure, and possible tectonic framework:
v. 7, p. 1123–1139. Academia Brasileira de Ciencias, Anais, v. 53, p. 361–373.
Eubank, R. T. and A. C. Makki, 1981, Structural geology of Hervé, F., and C. Mpodozis, 1990, Terrenos tectonoestratigrá-
the central Sumatra back-arc basin: Indonesian Petroleum ficos en la evolución geológica de los Andes Chilenos: una
Convention, Proceedings of the Tenth Annual Conven- revisión: Noveno Congreso Geológico Argentino, San
tion, p. 121–148. Juan, Actas 2, p. 319–323.
Eussler, A., 1970, Los reservorios petrolíferos de la Cuenca Keidel, J., 1925, Sobre el desarrollo paleogeográfico de las
del Golfo San Jorge: Primer Simposio de Recuperación grandes unidades geológicas de la Argentina: Anales de la
Secundaria de Petróleo y Gas, v. 1, p. 28. Sociedad Argentina de Estudios Geográficos, v. 4,
Fajing, C. H., and Z. Shulin, 1991, Tectonic evolution and p. 129–152.
hydrocarbon migration in the Guanhua basin, NE China: Laffite, G. A., and H. J. Villar, 1982, Poder reflector de la
Journal of Petroleum Geology, v. 14, p. 197–210. vitrinita y madurez térmica: aplicación en el sector NO de
Ferello, R., and P. Lesta, 1973, Acerca de la existencia de una la Cuenca del Golfo San Jorge: Primer Congreso de Hidro-
dorsal interior en el sector central de la Serranía de San carburos, Petróleo y Gas, Exploración, p. 171–182.
Bernardo (Chubut): Quinto Congreso Geológico Legarreta, L., M. A. Uliana, and M. A. Torres, 1990, Secuen-
Argentino, Actas, Tomo 5, p. 19–26. cias depositacionales cenozoicas de Patagonia central: sus
418 Peroni et al.

relaciones con las asociaciones de mamíferos terrestres y Shaw, R. D., M. A. Etheridge, and K. Lambeck, 1991, Devel-
episodios marinos epicontinentales. Evaluación prelim- opment of the Proterozoic to mid-Paleozoic, intracratonic
inar: Segundo Simposio sobre el Terciario de Chile, Amadeus basin in central Australia: a key to under-
Facultad de Ciencias de la Universidad de Concepcion, standing tectonic forces in plate interior: Tectonics, v. 10,
Actas, p. 135–176. p. 688–721.
Lesta, P. J., 1968, Estratigrafía de la Cuenca del Golfo San Jorge: Simpson, G. G., 1940, Review of the mammal-bearing
Terceras Jornadas Geológicas Argentinas, Actas 1, Tertiary of South America: American Philosophical
p. 251–289. Society Proceedings, v. 83, p. 649–709.
Lesta, P. J., and R. Ferello, 1972, Región extraandina del Skarmeta, J., 1976, Estratigrafía del Terciario sedimentario
Chubut y norte de Santa Cruz, in A. F. Leanza, ed., continental de la región central de la Provincia de Aisén,
Geología Regional Argentina: Academia Nacional de Chile: Asociación Geológica Argentina, Buenos Aires,
Ciencias, Córdoba, p. 601–654. Revista 31, p. 73–78.
Lesta, P. J., R. Ferello, and G. Chebli, 1980, Chubut Spalletti, L. A., and M. M. Mazzoni, 1979, Estratigrafía de la
extraandino, in J. C. M. Turner, coord., Segundo Simposio Formación Sarmiento en la Barranca Sur del Lago Colhue-
de Geología Regional Argentina: Academia Nacional de Huapi, Provincia del Chubut: Asociación Geológica
Ciencias, Córdoba, v. 2, p. 1307–1387. Argentina, Revista 24, p. 271–278.
Letouzey, J., 1990, Fault reactivation, inversion and fold- Talbot, M. R., 1988, The origins of lacustrine oil source rocks:
thrust belt, in J. Letouzey, ed., Petroleum and tectonics in evidence from the lakes of tropical Africa, in A. J. Fleet, K.
mobile belts: Editions Technip, Paris, p. 101–128. Kelts, and M. R. Talbot, eds., Lacustrine source rocks: The
Lovelock, P. E. R., 1984, A review of the tectonics of the Geological Society Special Publication 40, p. 29–44.
northern Middle East region: Geological Magazine, v. 121, Teruggi, M. E., and H. Rosetto, 1963, Petrología del Chubu-
p. 577–587. tiano del codo del Rio Senguerr: Boletín de Informaciones
Mattauer, M., P. Tapponier, and F. Proust, 1977, Sur les Petroleras, Buenos Aires, v. 354, p. 18–35.
mécanismes de formation des chaines intracontinentales. Ugarte, F. R., 1966, La cuenca compuesta carbonífera-jurásica
L’exemple des chaines atlasiques du Maroc: Societé de la Patagonia meridional: Universidad de la Patagonia,
Geólogique de France, Paris, v. 29, p. 521–526. San Juan Bosco, Comodoro Rivadavia, Ciencias Geológ-
Meconi, G. R., 1989, Estratigrafía y paleoambientes del Grupo icas, v. 2, p. 37–68.
Chubut en el codo del río Senguerr, Provincia del Chubut: Uliana, M. A., K. T. Biddle, D. W. Phelps, and D. Gust, 1985,
Master’s thesis, Universidad de Buenos Aires, Buenos Significado del vulcanismo y extensión mesojurásica en el
Aires, Argentina, 183 p. extremo meridional del Sudamérica: Asociación Geológica
Miller, H., 1976, El basamento de la provincia de Aysen Argentina, Revista 40, no. 4, p. 231–253.
(Chile) y sus correlaciones con las rocas pre-mesozoicas de Uliana, M. A., K. T. Biddle, and J. J. Cerdan, 1989, Mesozoic
la Patagonia argentina: Sexto Congreso Geológico extension and the formation of Argentine sedimentary
Argentino, Actas 1, p. 125–141. basins, in A. J. Tankard and H. R. Balkwill, eds., Exten-
Musacchio, E. A., 1981, Estratigrafía de la Sierra Pampa de sional tectonics and stratigraphy of the North Atlantic
Agnia en la región extrandina de la Provincia de Chubut, margins: AAPG Memoir 46, p. 599–614.
Argentina: Octavo Congreso Geológico Argentino, Actas Urien, C. M., and J. J. Zambrano, 1973, The geology of the
3, p. 343–357. basins of the Argentine continental margin and Malvinas
Oldow, J. S., A. W. Bally, and H. Ave Lallemant, 1990, Trans- Plateau, in A. E. M. Nairn and F. G. Stehli, eds., The Ocean
pression, orogenic float and lithosphere balance: Geology, Basins and Margins, v. I, The South Atlantic: Plenum
v. 18, p. 991–994. Press, p. 135–170.
Pittion, J. L., and J. Goudain, 1991, Source rocks and oil Van Niewenhuise, D. S., and A. R. Ormiston, 1989, A model
generation in the Austral basin, in John Wiley, ed., Source- for the origin of source-rich lacustrine facies, San Jorge
rock geology, 13th World Petroleum Congress, Topic 2, basin, Argentina: Primer Congreso Nacional de Explo-
p. 1–8. ración de Hidrocarburos, Tomo 2, p. 853–884.
Ramos, V. A., 1983, Evolución tectónica y metalogénesis de la Wildi, W., 1983, La chaine tello-rifaine (Algérie, Maroc,
Cordillera patagónica: Segundo Congreso Nacional de Tunisie): structure, stratigraphie et évolution du Trias au
Geología Económica, Actas 1, p. 107–124. Miocene: Revue Géologie Dynamique et de Géographie
Rapela, C. W., 1990, El magmatismo gondwánico y la Physique, Paris, v. 24, p. 197–201.
megafractura de Gastre: Décimo Primer Congreso Windhausen, A., 1924, Líneas generales de la constitución
Geológico Argentino, San Juan, Actas, Tomo I, p. 113–116. geológica de la región situada al oeste del Golfo de San
Rodrigo Gainza, L., O. Decastelli, M. Iñiguez Rodriguez, and Jorge: Boletín Academia Nacional de Ciencias, Córdoba, v.
J. Seiler, 1984, Análisis estratigráfico y oleogenético del 27, p. 167–320.
sector NO de la Cuenca del Golfo San Jorge: Noveno Yllañez, E. D., J. P. Di Lena, and H. G. Marchese, 1989, Evalu-
Congreso Geológico Argentino, Buenos Aires, Actas, ación geoquímica de petróleo y rocas generadoras en la
Tomo VII, p. 86–105. cuenca del Golfo San Jorge, Argentina: Primer Congreso
Roll, A., 1938, Estudio geológico de la zona sur del curso Nacional de Exploración de Hidrocarburos, Actas, Buenos
medio del Rio Deseado (Patagonia): Boletín de Informa- Aires, Tomo 1, p. 1127–1158.
ciones Pertroleras, Buenos Aires, v. 15, p. 3–60. Zaiyi, T., 1990, The formation and distribution of
Schull, T. 1988, Rift basins of interior Sudan: petroleum Mesozoic–Cenozoic sedimentary basins in China: Journal
exploration and discovery: AAPG Bulletin, v. 72, of Petroleum Geology, v. 13, p. 19–34.
p. 1128–1142. Ziegler, P. A., 1987, Late Cretaceous and Cenozoic intra-plate
Sciutto, J. C., 1981, Geología del codo del Rio Senguerr, compressional deformation in the Alpine foreland—a
Chubut, Argentina: Octavo Congreso Geológico geodynamic model: Tectonophysics, v. 137, p. 389–420.
Argentino, Buenos Aires, Actas, v. 3, p. 203–219.
Hydrocarbons in Inverted Segment of Andean Foreland: San Bernardo Belt, Central Patagonia 419

Authors’ Mailing Addresses

G. O. Peroni
A. G. Hegedus
J. Cerdan
L. Legarreta
M. A. Uliana
Astra C.A.P.S.A.
Exploración y Producción
Tucuman 744, Piso 7
1049 Buenos Aires
Argentina

G. Laffitte
Yacimientos Petroliferos Fiscales S.A.
Roque Sáenz Peña 777
1049 Buenos Aires
Argentina
Andean Basins

ILLIMANI of the Cordillera Real, viewed across Lake Titicaca, Bolivia, at


4200 m altitude, the highest navigable lake in the world. Illimani (6402 m)
consists of lower Paleozoic strata in discordant contact with a middle Tertiary
granitic batholith and is elevated along a west-vergent thrust sheet.

Edgar Ortiz, 1994, watercolor, 30 x 23 cm


Petroleum Geology of the Sub-Andean Basins of Peru

Jeremy M. P. Mathalone Manuel Montoya R.


PetroSantander Inc. CIA Consultora de Petroleo S.A.
Houston, Texas, U.S.A. Lima, Peru

Abstract

S ub-Andean Peru comprises the Marañon, Ucayali, and Madre de Dios basins which, together with three
subsidiary basins, cover an area of 370,000 km2. These basins extend considerable distances northward
into Ecuador and Colombia and southeastward into Bolivia. More than 5 billion bbl of recoverable oil have
been discovered in these basins, of which over 1 billion bbl of oil and almost 7 tcf of gas are in Peru.
The Tertiary foreland basins in front of the Eastern Cordillera are filled with up to 4 km of Tertiary molasse
sedimentary rocks. The basins are mainly of Miocene age and overlie older Paleozoic and Mesozoic depocen-
ters. Three major compressional episodes are recognized: a Middle Triassic event, an Early Cretaceous event
associated with major unconformities in some areas, and a regionally pervasive late Miocene–Pliocene
(Quechua III) event expressed in thrusting and compressional folding over most of sub-Andean Peru.
Two families of oils are differentiated in the Marañon basin, related to Permian and Cretaceous source
rocks. Three groups of oils in the Ucayali basin derive from Devonian, Carboniferous, Permian, and Triassic
sources. Oil samples in the Madre de Dios basin correlate to Devonian and Carboniferous shales. A variety of
trap types have been identified. The foreland can be divided into areas where preexisting faults have been
reversed by late Tertiary compression and flexural uplift, and areas unaffected by this deformation where
older, more subtle traps are important. To the west, the sub-Andean belt comprises regions where basement-
involved thrusts predominate and other areas characterized by thin-skinned thrusting. The Oriente-Marañon-
Ucayali basin complex has at least one large hydrocarbon accumulation in each trap type. The level of explor-
ation is low, and many areas are virtually unexplored.

Resumen

L a zona sub andina del Perú incluye las cuencas de antepaís de Marañón, Ucayali y Madre de Dios asi
como las cuencas de Santiago, Huallaga y Ene en el piedemonte. Conjuntamente estas cuencas cubren
unos 370 000 km2. Las cuencas se extienden hacia el norte en Ecuador y Colombia y hacia el sur en Bolivia. Mas
de 5 billones de barriles han sido descubiertos en estas cuencas, de los cuales un poco más de un billon de
barriles de aceite y casi 7 TCP de gas en Perú.
En frente de la Cordillera Oriental las cuencas de antepaís están rellenas por hasta 4 km de sedimentos
molásicos Terciarios. Estas cuencas de antepaís son principalmente de edad Miocena y cubren cuencas Pale-
ozóicas y Mesozóicas. Se reconocen tres episodios principales de deformaciones compresionales: un evento en
el Trias medio, un evento en el Cretáceo inferior asociado con una discordancia regional en ciertas áreas y un
evento de edad Mioceno tardio-Plioceno (Quechua III) expresado por pliegues y cabalgamientos los cuales se
observan regionalmente en el sub andino Peruano.
Se distinguen dos familias de aceites en la cuenca de Marañón que provienen respectivamente de rocas
fuentes Pérmicas y Cretácicas. En la cuenca Ucayali se reconocen tres tipos de aceites que derivan respectiva-
mente del Devónico, Carbonífero, Pérmico y Triásico. Unas muestras de petróleo de la cuenca Madre de Dios
corelacionan con unas arcillas del Devónico y del Carbonífero. Se identificaron varios tipos de trampas. La
zona de antepaís se puede dividir en dos; con un área donde unas fallas antiguas fueron invertidas durante la
compresión del Terciario superior con levantamiento de las estructuras y flexura de los sedimentos, y un área
que no ha sido afectado por esta ultima deformación y donde unas trampas más antiguas y más sútiles son
importantes. Hacia el oeste el cinturón del piedemonte incluye zonas donde los cabalgamientos involucran el
basamento y otras zonas donde los cabalgamientos no involucran el basamento y se horizontalizan en la
cobertura sedimentaria. La cuenca Marañón-Oriente tiene por lo menos una acumulación de hidrocarburos
importante en cada tipo de trampa. El nivel de exploración es bajo y varias áreas están virtualmente inexplo-
radas.

Mathalone, J. M. P., and Montoya R., M., 1995, Petroleum geology of the sub-Andean 423
basins of Peru, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of
South America: AAPG Memoir 62, p. 423–444.
424 Mathalone and Montoya

Putumayo Brazilian
Shield 80
Basin

70
Oriente COLOMBIA
ECUADOR
Basin
60

Maranon 50
BRAZILIAN

Number of Fields
Basin
A' SHIELD
Iquitos
40

PERU
Cushabatay
Santiago High BRAZIL
A Basin
30
2 km Base
Cretaceous Contour
Contaya 20
Bagua High Oil Field
Basin Gas Field
10
Basins 2
Huallaga Lines of 1
Basin section on Fig. 3 0
8° Pucallpa
Eastern B' Regional seismic 0 10 30 100 300 1000 3000
Cordillera line Fig. 4 Field Size (Million BBLS Oil)
E. Ucayali
Andes B Well location
Mountains Shira Basin for Fig. 10
High KEY AREA TOTAL NUMBER OF NUMBER OF SUCCESS
RESERVES WILDCAT DISCOVERIES RATE (%)
W. Ucayali Fitzcarrald (MMBBLS) WELLS
Basin High Putumayo 400 MM 67 22 33%
C'Madre (Colombia)
De Dios Maranon/Ucayali 1,350 MM* 113 29 26%
Basin BOLIVIA (Peru)

12° Lima Ene D' Oriente 3,300 MM 165 47 28%


Basin
Pacific Urubamba C (Ecuador)
5,050 MM 345 98 29%
High Eastern
Ocean Cordillera
0 100 200

km D
75° 71°
Figure 2—Field size distribution and discovery rates for
Figure 1—Foreland basins of Peru showing tectonic Ucayali-Marañon-Oriente-Putumayo basin complex. The
framework and distribution of oil and gas fields. Locations reserve total for the Marañon and Ucayali basins (1350
of Figures 3, 4, and 10 are shown. million bbl) includes 600 million bbl proven and probable
condensates estimated for the Camisea area.
INTRODUCTION
The sub-Andean foreland basins span the length of (BOPD). Exploration, in contrast, decreased in the late
the Andes from Venezuela in the north to the southern- 1970s as other operators were unable to emulate the
most tip of South America. All the major basins and success of Petroperu and Occidental. In the middle 1970s,
many of the smaller basins are hydrocarbon bearing, and Cities Service and Andes Petroleum recorded over 2000
all have been substantially deformed relatively recently. km of seismic data in the Madre de Dios basin and
The Marañon, Ucayali, and Madre de Dios are the drilled five dry holes; these are the only wells drilled in
principal foreland basins. Each has subsidiary basins, the Peruvian Madre de Dios. In 1979–1980, Shell signed
such as the Ene, Huallaga, and Santiago basins, where blocks 38 and 42 in the southernmost Ucayali basin.
they have been deformed and partitioned in the foothills Some seismic acquisition followed the drilling of the
of the Andean fold and thrust belt (Figure 1). Sepa well, which recovered only a small amount of
Despite the drilling of the first well (the Agua Caliente Carboniferous oil. The San Martin well, spudded in 1983,
field discovery) in 1938, exploration of the sub-Andean flowed 41 million cubic feet of gas per day (MMCFGD)
basins of Peru has mainly taken place from 10 to 25 years and 1626 bbl of condensate per day (BCD) from Creta-
ago. During this time, most of the 122 wildcat wells were ceous sandstones. The Cashiriari wildcat well drilled in
drilled and seismic coverage was extended to much but 1986 flowed at similar rates mainly from the same reser-
not all of the prospective basinal areas. The results of this voirs. Present proven reserve estimates are about 6.5 tcf
exploration effort and the large volume of geologic work gas and 400 million bbl of condensate. Mipaya 1X,
in the area, particularly by Petroperu and Occidental however, encountered modest amounts of gas in
who discovered most of the oil during this phase, Permian reservoirs, indicating that not all the major
indicate that a complex petroleum system exists, with surface anticlines in this area were charged with large
multiple hydrocarbon sources from lower Paleozoic to gas-condensate accumulations. Failing to reach an
Cretaceous. agreement to develop this resource, Shell ceased all
On the strength of these oil discoveries, a trans- exploration activities in 1988 and also relinquished the
Andean pipeline was completed in 1977 for an annual Madre de Dios basin foothills acreage where they had
production in excess of 100,000 bbl of oil per day recorded 500 km of seismic data.
Petroleum Geology of the Sub-Andean Basins of Peru 425

W E
A CAPIRONA PAVAYACU A'
0 SANTIAGO MARANON BASIN
BASIN
Depth - km

10

15
SW AGUA NE
B AGUAYTIA B'
CALIENTE
0 W. UCAYALI UCAYALI BASIN
Depth - km

5 Tertiary-Recent

10 Cretaceous

15 U - Trias - Jurassic
S N
C CASHIRIARI SAN MARTIN C'
Triassic Salt
0 S. UCAYALI BASIN
Depth - km

5 Upper Paleozoic

10 Lower Paleozoic

15 Pre - Ordovician
Basement
S N
D PARIAMANU D'
0 MADRE DE DIOS BASIN
Depth - km

10
0 50 100 km
15

Figure 3—Geologic cross sections through the foreland basins of Peru showing the composite form of pre-Cretaceous
extensional basins that have been modified by younger encroachment of the Cordilleran fold and thrust belt. The climax of
Andean deformation and foreland basin subsidence occurred in the late Cenozoic. See Figure 1 for locations (A–A' through
D–D'). (Modified in part from Petroperu, 1989.)

By the end of the 1980s, there was little exploration of basement fabric was established by accretion of conti-
the sub-Andean basins in Peru, with only Petroperu and nental fragments or terranes to the edge of the Brazilian
Occidental being active. Mobil Exploration signed blocks shield. At the end of the Precambrian and beginning of
covering most of the Huallaga basin in 1989, recording the Cambrian, the Brasiliano tectonic event created the
seismic data prior to drilling a dry well. Production from structural framework that has been repeatedly reacti-
the Peruvian sub-Andean fields also dropped, from vated throughout the Phanerozoic (Sempere, 1995). Early
128,000 BOPD in 1980 to 77,000 BOPD in 1992, reflecting Paleozoic sediments were deposited on a fault-controlled
the lack of new discoveries. During the 1990s, however, passive margin adjacent to a collision zone (Figure 3).
in response to more favorable contract terms, a number Subduction-related granitoids were emplaced in
of companies have negotiated contracts and exploration southern Peru in Late Ordovician–Early Devonian time
activity is again increasing. Figure 2 summarizes the field (Mukasa and Henry, 1990). Late Devonian–Early
size distribution for the Oriente, Marañon, and Ucayali Carboniferous deformation is reflected in a regional
basins. unconformity resulting from the Chanic orogeny of
This paper presents a simple, integrated review of the northwestern Argentina (equivalent to the Caledonian).
petroleum geology of the sub-Andean basins of Peru. Tertiary orogenesis built the Andean fold and thrust belt
Some data from the Oriente basin of Ecuador are included as well as the flexural foreland basins. Compressive reac-
because it is completely contiguous with, although tivation and inversion of the older basement-involved
surprisingly different from, the Peruvian Marañon basin. fault system dissected and partitioned the foreland basin
into several depocenters (Figure 3).
The Marañon basin is part of a much larger structural
BASIN SETTING basin that includes the Oriente basin of Ecuador and the
Putumayo basin of southern Colombia (Figure 1). The
The Peruvian margin of South America preserves a whole province is a typically asymmetric foreland basin
long history of deformation and basin formation. The covering 320,000 km2. The western margin of the basin
426 Mathalone and Montoya

YANAYACU OILFIELD
SW NE
SECS
TWT

1.0

Top Cretaceous
2.0

Base Cret.
3.0
Base Pucara Fm
4.0
Top Carb.
?
Basement
0 25 50 km

Figure 4—Regional seismic cross section, Marañon basin, showing Paleozoic extensional fault and half-graben, westward-
thickening Mesozoic succession, basement-involved thrusting to the southwest, and a thick Tertiary foreland basin succes-
sion. Tertiary compression has reactivated the Paleozoic normal faults in the reverse sense and created local forced folds.
See Figure 1 for location.

consists of the 50–200-km-wide sub-Andean belt of Cashiriari and San Martin gas fields. The Ucayali basin is
thrusted Mesozoic sedimentary rocks that abut the large separated from the adjacent Ene basin of the foothills by
Santiago, Huallaga, and Ene basins (Figures 1, 3). The the reactivated Shira high (Figures 1, 3).
Marañon and its subsidiary basins contain up to 4000 m The Madre de Dios basin, covering 120,000 km2, is a
of Tertiary molasse deposits that were shed from the southward-dipping foreland basin bounded on the south
Cordillera rising to the west. The extreme western part of by a 40–70-km-wide zone of detached thrusts that form
the basin is dominated by a series of back-thrusts that the sub-Andean foothills. More than half of the basin lies
isolated the Santiago basin (Rodriguez, 1982). Under- in Bolivia where it is bounded on the east by the Madidi
lying the Tertiary molasse fill is a westward-thickening high (Figure 1). The Madre de Dios basin is notable in
Cretaceous clastic section that contains much of the that the foreland monocline appears to be virtually un-
source and most of the proven oil reserves as well as an deformed, although there are large gaps in seismic
older Triassic–Jurassic wedge. A suite of Paleozoic rift coverage. The succession consists of Devonian–Permian
basins have suffered minor inversion in the eastern marine shales and carbonates, thin Cretaceous conti-
foreland. The geometry of this basin fill and the local nental strata, and the ubiquitous Tertiary cover sequence.
inversion are shown in the seismic profile of Figure 4. The thin Devonian–Cretaceous argillaceous section
In contrast to the broad simplicity of the Marañon exceeds 1500 m in the depocenter near the Bolivian
basin, the Ucayali basin is formed into a series of west- border. A major unconformity separates the Devonian
dipping asymmetric lows by massive basement-involved and Carboniferous rocks. Only eight wells have been
east-verging thrusts (Figure 1). There are two principal drilled in this remote basin (including three in Bolivia),
depocenters, the West Ucayali basin (or Pachitea basin) where exploration focuses on Paleozoic intervals in
and the East Ucayali basin (Figure 3), which coalesce to contrast to the basins farther north where the Mesozoic
the north. The Ucayali basin is bounded on the north by section is the major target.
the Contaya and Cushabatay uplifts, on the west by the
overthrusted Andean range, and on the south by the
Fitzcarrald high that constricts the basin. This high is STRATIGRAPHY
associated with a northeast-trending normal fault system
of Paleozoic origin and with an abrupt change in the Paleozoic
trend of the Andean Cordillera and associated faults.
The sedimentary fill of the Ucayali basin is similar to The sedimentary section of sub-Andean Peru ranges
the Marañon basin, comprising up to 3000 m of Tertiary from Ordovician to Recent in age (Figure 5). Although
continental molasse clastics overlying westward-thick- the present geometry of the basin reflects late Tertiary
ening marine-dominated wedges of Triassic, Jurassic, deformation, most of the succession is of significance to
and Cretaceous age. Unlike the basins farther north, the hydrocarbon exploration (Figure 6). Economic basement
Paleozoic section in the Ucayali basin is marked by for the petroleum industry in eastern Peru comprises
conspicuous thickening of the Devonian Cabanillas upper Precambrian metamorphic formations with a
Formation to over 800 m in the Camisea area. The dominant northwest-southeast grain above the lower
dominant structural form of the Ucayali basin is major Proterozoic cratonic rocks of the Arequipa massif, under
basement-involved thrusting, but in the extreme south, the present Eastern Cordillera and sub-Andean basins to
there is a zone of thin-skinned thrusting and folding that the west and the Brazilian shield to the east. The earliest
is expressed in the anticlines that form the giant Paleozoic rocks deposited over this basement are the
Petroleum Geology of the Sub-Andean Basins of Peru 427

MARANON UCAYALI MADRE DE MAJOR


BASIN BASIN DIOS BASIN TECTONIC Brazilian
EVENTS Shield
PLIO CORRIENTES/MARANON
QUECHUA III
Maranon/Oriente
Basin
TERTIARY

MIO
IPURURO PEBAS QUECHUA II
QUECHUA I
CHAMBIRA COLOMBIA
OLIG
POZO ECUADOR
EOC INCAIC
PAL
YAHUARANGO BRAZIL
PERU
SCU/HU/CASABLANCA PERUVIAN
UPPER

VIVIAN VIVIAN Iquitos


U. Jurassic
CRETACEOUS

CHONTA CHONTA 4° L. Jurassic


AGUA CALIENTE U. Triassic
LOWER

RAYA (ESPERANZA) ORIENTE L. Triassic


GROUP U. Permian
CUSHABATAY
NEVADAN L. Permian
SARAYAQUILLO GROUP Carboniferous
JURASSIC
Devonian
to Ordovician
PUCARA GROUP
Basement
TRIASSIC
TRIASSIC
MITU GROUP EVENT
PERMIAN ENE LATE 8° Huallaga Pucallpa
COPACABANA HERCYNIAN
TARMA Basin Ucayali
CARBONI-
FEROUS AMBO Basin
CHANIC
DEVONIAN CABANILLAS
Andes
SILURIAN ANANEA/SAN GABAN TACONIAN Mountains
ORDOVICIAN CONTAYA UPLIFT
CAMBRIAN
PRECAMB
BASEMENT

Significant Hydrocarbon Significant Madre BOLIVIA


Occurrence Source Rock De Dios
Lima
12° Basin
Ene
Pacific Basin
Ocean
Figure 5—Comparative stratigraphic column for composite 0 100 200
Marañon, Ucayali, and Madre de Dios basins. Quechua III
km
was a major diastrophism. Peruvian tectonism is only 75° 71°
recorded in Cordillera outcrops. Equivalent Cretaceous
formation names in Ecuador are noted in Figure 10. Figure 6—Cretaceous subcrop map.

organic-rich shales and arenites of the Ordovician marine subordinate tuffs. After a period of nondeposition in the
Contaya Group. Less than 1000 m are preserved, Namurian, the Westphalian was marked by transgres-
although considerably greater thicknesses are thought to sion and deposition of the Tarma Formation, a thin cover
have been originally deposited in southern Peru. The of sandstones, carbonates, and local tuffs. The lack of
Ordovician thins eastward onto the Brazilian shield and coarse clastics reflects a low-relief source area. These
is absent over some regional highs such as the Contaya terrigenous strata are generally followed unconformably
and Shira platforms, indicating the antiquity of these by the thicker carbonate Copacabana Group of West-
structural elements. phalian–Early Permian age. These are typically fossilif-
Little Silurian deposition or erosion is thought to have erous platform carbonates that grade laterally into evap-
occurred; fine-grained marine clastics of the Devonian oritic facies. The Copacabana limestones covered most of
Cabanillas Formation overlie the Ordovician with only sub-Andean Peru, except over the Contaya arch (Figure
minor disconformity. Glacial sedimentary rocks in 1) where the Cretaceous erosively overlies lower
Bolivia indicate possible eustatic sea level fall at this time. Paleozoic rocks (Eduardo, 1991).
The Devonian Cabanillas Formation is thickest (over The Copacabana Group is succeeded conformably by
1500 m) beneath the Madre De Dios basin near the the Ene Formation, comprising up to 600 m of black
Bolivian border. This formation is mainly shale prone organic-rich shales and dolomites with minor sandstones
with local development of coarse-grained clastics in (Figure 7). This is a regressive hypersaline sequence that
small coarsening-upward deltaic sequences. Deforma- originally covered most of the area prior to Triassic
tion during the Late Devonian and Early Carboniferous erosion. All of these formations onlap the western and
is expressed in local low-grade metamorphism and a southern flanks of the Contaya high, which is believed to
stratigraphic hiatus in the Marañon and Ucayali basins have been emergent throughout the late Paleozoic.
(Figure 5). Toward the end of the Permian and earliest Triassic, the
Carboniferous–Lower Permian sediments were entire area was uplifted and continental clastics and
deposited in a basin with a northwest-oriented depo-axis volcaniclastics of the Mitu Formation were deposited in
broadly following the trend now occupied by the Eastern small, fault-bounded extensional basins. The Mitu
Cordillera. The lowest Carboniferous rocks are the Formation is one of the few pre-Andean units that clearly
fluvial clastics of the Dinantian Ambo Group, shows thickening into faults, implying syndepositional
comprising conglomerates, sandstones, shales, coals, and faulting.
428 Mathalone and Montoya

Brazilian Marañon and Ucayali basins. The depocenter appears to


Shield
have been west of the present basins, the preserved
Maranon/Oriente
Basin Pucara section forming a spatially restricted westward-
COLOMBIA
thickening wedge of sediments (Figures 6, 8B).
ECUADOR Deposition in this Late Triassic basin started with
BRAZIL
restricted halites and anhydrites of the basal Pucara
Group, a locally thick and mobile layer that accommo-
dated later thin-skinned detachment. These evaporites
Iquitos
are succeeded by widespread platform limestones,
4° PERU
argillaceous limestones, and shales that are over 1000 m
thick in outcrops in the Cordillera and suggest a paleo-
Good Oil Correlation bathymetry that shallowed toward the east. The typical
with Ene Source
organic richness of these sedimentary rocks testifies to
Partial Oil Correlation
with Ene Source the restricted nature of the basin at that time.
Early movement of the Pucara halites together with
Outcrop sample
TOC >2.5% regional uplift along the western margin of the basin
Outcrop sample
affected deposition of the regressive Upper Jurassic
8° Pucallpa
TOC <2.5% Sarayaquillo Formation. This suite of continental sand-
Ucayali stones, conglomerates, and red beds blanketed the
Andes Basin Pucara Group but, unlike the underlying marine
Mountains
deposits, they thin westward. The boundary between the
Sarayaquillo and the underlying Pucara Group is
Madre
De Dios generally unconformable, reflecting onset of earliest
Basin BOLIVIA Cretaceous (Nevadan) tectonism and indicating early salt
12° Lima
Ene
movement. The close of the Jurassic, like many basins
Basin worldwide, is marked by a regional Nevadan unconfor-
Pacific
Ocean mity, commonly the only unconformity visible on
0 100 200
seismic data (Figure 4). A hiatus of about 20 m.y. is
km
inferred between the top of the Upper Jurassic Saraya-
75° 71°
quillo Formation and the overlying Cretaceous succes-
Figure 7—Permian Ene formation distribution showing oil sion.
to source correlation. The Cretaceous comprises a westward-thickening
wedge of fluvial and marginal marine clastics that occur
in five formations (Figures 5, 9, 10): two shale-prone units
Mesozoic sandwiched between three sandstone sequences. Deposi-
tion is attributed to tectonic as well as eustatic processes
Early Triassic uplift and erosion were contempora- as suggested by the persistence of the stratigraphy for at
neous with Mitu deposition and were associated with least 1500 km (Figure 10). The Cushabatay Formation
voluminous igneous activity in the area of the present represents the initial marine transgressive sandstones of
Cordillera. The Triassic event, colloquially referred to as the Cretaceous. They are thickly bedded, medium- to
the Jurua orogeny, was one of the most significant coarse-grained cross-bedded quartz arenites that are
erosional events in the geologic history of sub-Andean interbedded with subordinate shales containing
Peru. Throughout much of the Marañon and Ucayali abundant plant detritus. These sandstones were
basins, major faulting and erosion alternately preserved deposited in a delta plain to marginal marine setting;
and eroded thick Paleozoic sections (Figure 8A). The pre- they form good hydrocarbon reservoirs.
Andean terrane of the Marañon and Ucayali basins thus Overlying the Cushabatay sandstones is the Albian
comprises a suite of Paleozoic rifts and interbasin highs Raya Formation (Figure 5), consisting of up to 300 m of
without cover deposits (Figure 8A). This event marks the organic-rich marine shales interbedded with minor silt-
transition to the Andean orogeny. Little faulting or stones, fine-grained sandstones, and subordinate sandy
erosion is apparent in the Madre de Dios basin, where limestones. To the west, the shales are rich in organic
relatively uneroded pre-Andean sedimentary rocks are material and were deposited in a shallow restricted
conformably overlain by the Cretaceous succession. An marine environment. Toward the Cretaceous deposi-
exception is the Cenozoic foreland basin, where Creta- tional edge in the east, the Raya Formation grades into
ceous sediments unconformably overlie progressively less organic-rich littoral shales. The Raya Formation
older units down to the Devonian in a southward wedges out toward the south (Figure 10), where conti-
direction. However, it is believed that this erosion is of nental conditions prevailed over much of the Madre de
Early Cretaceous age (Figures 3, 8A). Dios basin throughout the Cretaceous.
Subsidence and marine transgression continued in the The middle Cretaceous Agua Caliente Formation
Late Triassic with deposition of the Upper Triassic– (Figure 10) comprises coarse-grained, large-scale cross-
Lower Jurassic Pucara Group. This succession was bedded sandstones. These are interbedded with fine-
deposited in a widespread low covering the western grained sandstones, siltstones, and black shales with
Petroleum Geology of the Sub-Andean Basins of Peru 429

Brazilian Brazilian
Shield Shield
Maranon/Oriente Maranon/Oriente
Basin Basin
COLOMBIA COLOMBIA
ECUADOR ECUADOR
BRAZIL

BRAZIL
PERU
Iquitos Iquitos
PERU

75
4° L. Triassic 4°

0
U. Permian Correlation of
L. Permian oil with Pucara

250
Carboniferous group source

500
Devonian Pucara Outcrop
to Ordovician
Outcrop sample
Basement TOC >2.5%

Thrusted Outcrop Sample


Outcrop TOC <2.5%
8° 8°
Pucallpa
Huallaga Ucayali 500 Pucallpa
C.I. = 250m Gross Thickness
Basin Basin Ucayali

250
Basin
Andes
Mountains Andes
Mountains Madre
De Dios
Basin
Madre
De Dios BOLIVIA
Basin BOLIVIA
Lima Ene Lima Ene
12° A Basin B Basin
12°

Pacific Pacific
Ocean Ocean
0 100 200 0 100 200

km km

75° 71° 75° 71°

Figure 8—Triassic basin. (A) Middle Triassic subcrop map. (B) Isopach of Triassic–Jurassic Pucara Group and total organic
carbon (TOC). (TOC largely after Touzett and Sanz, 1985.)

abundant plant remains. The sandstones are up to 350 m section consists of terrestrial sedimentary rocks.
thick and generally have favorable reservoir characteris- The Upper Cretaceous Vivian Formation is the most
tics. Like the rest of the Cretaceous stratigraphy, these important reservoir zone of the Peruvian sub-Andean
clastics were largely sourced from the Brazilian shield. basins (Figure 11B). It is generally a single sandstone
The uppermost Albian–lowermost Turonian Agua body, but thickens into a series of sandstones as much as
Caliente Formation grades from fluvial in the southern 150 m thick in the northeastern Marañon basin. In the
area to estuarine and marginal marine in the north. Bio- Oriente basin of Ecuador, its counterpart, the Napo M-1
stratigraphic evidence suggests a depositional hiatus sandstone, is widely eroded due to post-Cretaceous
within this formation spanning much of the Cenomanian uplift.
(the so-called Mochica event). The depositional environment of the Vivian
The greatest Cretaceous transgression occurred Formation varies over the area. It is fluvial in the north-
during the Coniacian and early Santonian when the eastern Marañon basin, the eastern part of the Ucayali
shale-prone Chonta Formation was deposited (Figure basin, and the entire Madre de Dios basin, but has
11A). The lower part of the Chonta Formation is marginal marine affinities to the west. A large number of
commonly arenaceous with nearshore glauconitic sand- wells were drilled by Occidental in the northern part of
stones, up to 3 m thick, interbedded with ripple cross- the Marañon basin adjacent to the Ecuadorian border. In
laminated mudstones. These bar and shoal deposits this area, the lower Vivian sandstones are clean, cross
grade upward into dark organic-rich inner shelf shales. bedded, and were deposited as elongate bodies with
Although present over most of sub-Andean Peru (Figure paleocurrents directed toward the southwest (Augusto et
10), the shales are best developed in the northern part of al., 1990). The sparcity of carbonaceous detritus indicates
the Marañon basin and its continuation in the Ecuado- an arid alluvial plain. The upper sandstones locally form
rian Oriente basin (Napo Formation), as shown by the marginal marine bar and channel sandstones. Overall,
organic carbon distribution (Figure 11A). The formation the Campanian Vivian Formation was deposited during
thins and becomes sandier toward the east and south a major regression following the Chonta highstand.
concomitant with shallowing. Farther south, in the South Most of the Maastrichtian interval was nondeposi-
Ucayali and Madre De Dios basins, the entire Cretaceous tional in the Madre de Dios basin, but is represented in
430 Mathalone and Montoya

Brazilian 600 Brazilian


Shield Shield

-2.0
Maranon/Oriente Maranon/Oriente
ECUADOR Basin Basin
COLOMBIA COLOMBIA
-3.0

20
ECUADOR
-4.0

0
-5.0

400
-60

800
Iquitos Iquitos
4° PERU 4°
PERU
1200

16
BRAZIL BRAZIL

00
-5.0 Foothills
-4.0 basin with
complex thrusting C.I. = 200 m
and folding
Santiago
Basin C.I. = 1 Km
.0
-1
8° Pucallpa Pucallpa 8°
Huallaga
Basin Ucayali Basin Ucayali
-2.0

Basin
-1.0

Andes
Mountains Andes
Mountains
Madre
Sub-Andean De Dios
Thrust Belt -1.0 Basin

400
Madre BOLIVIA
BOLIVIA
De Dios
-2.0 Basin
12° A Lima Ene
Basin -3.0 B Lima Ene 12°
Basin
Pacific -4.0
Pacific
Ocean Ocean
0 100 200
0 100 200

km
km
75° 71° 75° 71°
Figure 9—Cretaceous basin. (A) Depth to base of Cretaceous (partially after Salas, 1991). (B) Isopach of Cretaceous basin fill.

the Ucayali basin by three thin, terrigenous clastic units and subordinate sandstones. The unit is up to 300 m
with a combined thickness generally less than 250 m. The thick in the extreme west, but considerably thinner
lowest unit, the Cachiyacu Formation, represents a minor elsewhere. The overlying Pozo Formation is a sandstone
marine transgression and deposition of littoral to and tuff interval that is covered with thin shales attrib-
estuarine dark gray shales and interbedded, fine-grained uted to marine incursion. This upper Eocene–lower
sandstones. These shales form a good seal. The marginal Oligocene unit is ascribed to increased erosion of the
marine Cachiyacu Formation grades upward into red rising Cordillera and concomitant sag of the flexural
floodplain mudstones of the Huchpacayu Formation. basin due to the applied orogenic load; this deformation
The Cretaceous sequence of the Ucayali basin is capped is referred to as the Incaic event.
with thin marginal marine sandstones and shales of the The Oligocene–middle Miocene is represented by the
Casablanca Formation. These Maastrichtian clastics are Chambira Formation composed of thick, red shales with
not documented in most of the Marañon basin, where it locally interbedded alluvial plain sandstones. Typical of
is believed that they were eroded prior to the Tertiary. the other Tertiary molasse units, the Chambira thickens
westward into the Andean foreland basin. The upper
Tertiary Miocene Pebas Formation comprises mainly red bed
shales with local evaporites, reflecting tectonic quies-
The foreland basins of Peru contain a molasse wedge cence before the late Miocene–Pliocene Quechua phase
up to 4000 m thick that was deposited in front of the III deformation. The Pliocene is dominated by coarse-
encroaching Andean fold and thrust belt, mainly in late grained clastics that were deposited in response to this
Tertiary time. These deposits are of little commercial deformation.
interest, except for their sealing capabilities. The rapid
deposition of these molasse sediments and the formation
or modification of structural traps affects the generation, TECTONIC FRAMEWORK
migration, and entrapment of petroleum in earlier
deposits. The processes that formed the sub-Andean basins of
The Paleocene Yahuarango Formation, which Peru were related to evolution of the plate margin. An
conformably overlies the Cretaceous succession, Atlantic-type passive margin characterized the
comprises a westward-thickening wedge of red shales Cambrian–Early Ordovician continent edge. In western
Petroleum Geology of the Sub-Andean Basins of Peru 431

Figure 10—Cretaceous stratigraphic correlation. (The Ecuadorian example is after Canfield, 1991.)

Bolivia, the Puna trough accumulated over 10,000 m of a depositional hiatus (Figure 5).
flysch deposits (Dalmayrac et al., 1980). The earliest Extensional faulting occurred during the Carbonif-
collision is dated as Middle Ordovician with magmatic erous and Early Permian and was followed by minor
arc accretion, including the Puna eruptive belt of Mendez middle Permian (late Hercynian) compression. These are
et al. (1972). Subduction-related strike-slip deformation attributed to the transtension and transpression
and associated transtensional and transpressional condi- described by Sempere (1995) and have been overprinted
tions controlled basin formation until the Triassic by younger Andean deformation. The Paleozoic succes-
(Sempere, 1995). The modern subduction system is dated sion beneath most of the sub-Andean basins is largely
as Early Jurassic as the continental platform was undeformed. The Late Permian and Early Triassic were
subjected to an episode of extension before the opening characterized by extensional faulting with extensive
of the Atlantic Ocean. Conspicuous tectonic events at the magmatism in the present Andean region. A Middle
beginning of the Cretaceous and late Miocene (Quechua Triassic event is recorded in much of the Marañon and
III) reflect changes in plate convergence and stress fields. Ucayali basins. The subcrop of the units below the
The compressional tectonism resulted in inversion of Pucara Formation (Figure 8A) shows the patchy preser-
previous depocenters and eastward encroachment of the vation of Paleozoic rocks often in half-grabens at this
Andean fold and thrust belt. This late Tertiary deforma- unconformity.
tion created many of the structural traps and led to The Triassic–Jurassic Pucara Formation was deposited
source rock maturation and hydrocarbon generation and in a marine basin with a depocenter located where the
migration. Eastern Cordillera is now. The growth of a regional high
The earliest regionally identified deformational event during the Late Jurassic is suggested by the distribution
in sub-Andean Peru occurred during the Late Devonian of the Sarayaquillo Formation, which is restricted mostly
and earliest Carboniferous; this is referred to the Eoher- to the Marañon and Ucayali basins. Earliest Cretaceous
cynian and is equivalent to the Chanic orogeny of Bolivia tectonism is interpreted (Figure 4) on the basis of local
and northern Argentina. Deformation was intraconti- unconformities.
nental, resulting from contraction between the Arequipa A Late Cretaceous compressional event is ascribed to
massif to the west and the Brazilian shield. Compres- low-angle subduction of the Pacific plate beneath the
sional structures include northwest-oriented folds, Peruvian margin. Compressional structuring occurred at
cleavage, and metamorphism (Mégard, 1978; Martinez, this time throughout the Andean region, including the
1980). Over the thicker continental crust, there was little Altiplano (Sempere, 1995). However, the Marañon and
deformation. However, the diastrophism is expressed in Ucayali basins went largely unscathed because they were
432 Mathalone and Montoya

Brazilian Brazilian
Maranon/Oriente Shield Shield
Basin Maranon/Oriente
ECUADOR Basin
COLOMBIA BRAZIL ECUADOR
3.0

Er COLOMBIA
os BRAZIL
io

20
40
na

0
2.0 FORESTAL-1 CORRIENTES 12 FORESTAL-1 lE
FORESTAL-1 dg
2.0
1.0

1.0

60
e

1 0
12 00
1.0

14 0
9000

0
VIVIAN FORMATION
CORRIENTES 12 CORRIENTES 12

Iquitos Iquitos
4° PERU

DEPTH (FT)
PERU 9000 4°

0
0.5

16
?
BOLIVIA
BOLIVIA

9500 100

CHONTA FORMATION
C.I. 20m
VIVIAN
0•2-1•5% SP ILD
9500 Ucayali THICKNESS
0•2-1•1% Basin
100

80
8° Pucallpa 8°
Pucallpa Huallaga
Basin
40 60
20
Andes Ucayali SP ILD
Andes
Mountains Basin Mountains Madre
DECREASING Madre 10 0 De Dios
MARINE
De Dios 2 0
INFLUENCE, INCREASING 4 0 Basin
COARSE GRAINED Basin 6 0
CONTENT 8
BOLIVIA
BOLIVIA 80

60
12° A Lima Ene
Basin B Lima Ene
Basin
12°
Pacific Pacific 40
Ocean Correlation C.I. 0.5% Ocean
of oil with AVERAGE
0 100 200 0 100 200
Cretaceous TOC
source
km km

75° 71° 75° 71°

Figure 11—(A) Upper Cretaceous Chonta Formation showing TOC distribution and oil to source correlation. (B) Isopach map
of the Upper Cretaceous Vivian Formation and comparison of lithostratigraphy between the Forestal-1 and Corrientes wells,
which reflects basin dynamics.

on the fringe of this depositional tract. This tectonism is more recent Cushabatay high to form the saddle
believed to have been the start of Andean deformation between the Ucayali and Marañon basins (Figure 1). This
and mountain building. Sempere (1995) dates this event thrusting is associated with a major sinustral wrench
as late Turonian or early Coniacian. zone that trends WNW-ESE (Vernet and Xavier, 1990).
The Incaic period of compressional deformation took The Ene foothills basin was further separated from the
place in the early–middle Eocene and is thought to have Ucayali basin by thrust reactivation of the preexisting
been responsible for much of the shortening in the Shira high. All the foothills basins (Santiago, Huallaga,
central Andes. Like the previous Peruvian phase, and Ene) continued to develop as piggy-back basins and
however, there was no marked deformation in the sub- were intensely deformed along with the western margins
Andean basins. Instead there was regional flexural of the principal foreland basins.
tilting, onlap, and renewed structural growth. The Marañon basin was generally less deformed by
During the Miocene, three compressional events Quechua III movements than the Ucayali basin; the latter
ascribed to Quechua tectonism built the Andes into the is surrounded by major basement-involved thrusting
modern mountain belt (Mégard, 1984; Sempere, 1995). and inversion structures. The eastern foreland of the
There is no evidence in the sub-Andean basins (e.g., Marañon basin was reactivated by this deformation.
intra-Tertiary angular unconformities or onlap onto the Many listric normal extensional faults were inverted. The
flanks of folds) for the first two phases, but halotectonic western edge of the Marañon basin, although well
and halokinetic structuring may have been a response. faulted in the Paleozoic, did not suffer pronounced
The late Miocene Quechua III event, however, pro- inversion or structural telescoping, except for subtle
foundly affected most of sub-Andean Peru, as evidenced folding and forced folding above older structures.
by the present form of the foreland basins. A large set of Quechua III deformation persisted locally through the
salt-facilitated thrusts and back-thrusts (Figure 3) Pliocene.
separated the Santiago and Huallaga basins from the This late Tertiary or Quechua tectonism was respon-
main Marañon basin (Rodriguez, 1982). Farther south, sible for formation of hydrocarbon traps, including
the Contaya high was reactivated and thrust with the thrust anticlines, inverted normal fault structures, subtle
Petroleum Geology of the Sub-Andean Basins of Peru 433

folding, forced folds, and structural enhancement of minor reserves in sandstones of the Permian Ene
stratigraphic reservoirs and seals. Figure 12 shows the Formation. The Sepa-1X well in the southern part of the
seismic expression of these types of structures and the Ucayali basin recovered a few gallons of 33˚ API oil from
way they have contributed to a variety of oil and gas the Upper Carboniferous Copacabana limestone. Farther
fields. north in the eastern Ucayali basin, oil shows have been
reported from the La Colpa well. At the other end of the
basin, oil seeps between the northernmost Ucayali basin
PETROLEUM GEOLOGY and Huallaga basin are believed to be sourced from the
Triassic Pucara or older rocks.
Oil and gas seeps occur along the eastern edge of the Oil and gas seeps have been reported from the
Andes from Ecuador to northwestern Argentina. Oil Andean foothills adjacent to the Madre de Dios basin,
seeps in the Ucayali basin led to the early discovery of and a small oil accumulation apparently occurs in
the Maquia and Agua Caliente oil fields. Many oil seeps Devonian sandstones in the Bolivian part of this basin.
from Cretaceous and Tertiary sedimentary rocks occur in The Pariamanu well recovered a small amount of 44–53º
the Santiago basin, reflecting compressional and API oil from Carboniferous sandstones, while the Puerto
transtensional tectonism as well as the sand-prone Primo well recovered a few barrels of 42˚ API oil from
Tertiary basin fill. A suite of surface seeps between the Devonian sandstone. Like the southern Ucayali basin,
Huallaga and Marañon basins are associated with salt the occurrence of oil and gas indicates a Paleozoic origin
diapirs. for hydrocarbons in the Madre de Dios basin.
There have been 122 wildcat wells drilled in the sub-
Andean basins of Peru, mainly in the last 25 years. These Reservoir and Seal
have resulted in the discovery of 740 million bbl of recov-
erable oil in the Marañon basin and about 400 million bbl Almost all of the producible hydrocarbons discovered
of recoverable condensate liquids (7 tcf of gas) in the in the sub-Andean foreland basin are reservoired in
Ucayali basin. Potential reserves (resource endowment) Cretaceous clastics. Small amounts of oil have leaked
are considerably greater, including condensates from the into basal Tertiary sedimentary rocks of the Marañon
Shell Cashiriari and San Martin discoveries in the basin, and gas is trapped in sandstones of the Permian
Ucayali basin. Ene Formation in the southern Ucayali basin. However,
Oil in the Marañon basin is reservoired principally in the Cretaceous succession remains the focus of
terrigenous clastics of Cretaceous age, with a subordinate petroleum exploration in the Marañon and Ucayali
amount in Tertiary reservoirs. These oils are usually basins.
undersaturated with respect to gas, have formation Devonian sandstones of the Cabanillas Formation are
volume factors typically less than 1.1, and a range of enveloped in potential source rocks in the Madre de Dios
gravities from 10˚ to 42˚ API (see Appendix Tables A1 basin. A small discovery of oil reservoired in these
and A2). More than half of this crude oil (450 million bbl) clastics has been made in the Bolivian Madre de Dios
is reservoired in the Upper Cretaceous Vivian Formation. basin. Most of these sandstones are discontinuous and
Nearly all of the remaining 290 million bbl is reservoired generally of poor reservoir quality. Carboniferous sand-
in the Cretaceous Chonta Formation immediately below stones in the basal Tarma and upper Ambo formations
the Vivian Formation (Figure 5). also have an irregular distribution in the Ucayali and
The distribution of hydrocarbons in the Ucayali basin Marañon basins. They are fine-grained to conglomeratic
is different from that of the Marañon, being dominated sandstones with electric log porosity values exceeding
by gas. The giant Cashiriari and San Martin discoveries 15%. These reservoirs are overlain by Carboniferous–
made by Shell in the mid-1980s comprise over 7 tcf of gas Lower Permian Copacabana limestones. These platform
and 400 million bbl of condensate; probable reserves are carbonates are generally impermeable, but are suscep-
believed to be much larger. These fields, as well as the tible to compressional fracturing and are believed to
smaller Mipaya field, are located in the southern part of have doubtful seal integrity.
the Ucayali basin. The Copacabana carbonates are overlain by shales of
The small Aquaytia gas-condensate field (250 bcf) was the Permian Ene Formation, which has effective sealing
discovered in the western Ucayali basin. Some small oil properties. A shallow marine sandstone unit up to 50 m
accumulations have also been discovered in the Ucayali, thick occurs in this formation over much of the Ucayali
such as the Maquia light oil field, in a tract of oil-bearing and Ene basins. Significant amounts of gas are trapped in
compressional structures adjacent to the Contaya high this sandstone unit in the Mipaya and Cashiriari fields.
(Figure 1). The first oil discovery in the sub-Andean Although reservoir quality is poor to moderate, this
foreland basin was Agua Caliente (14.7 million bbl of 44˚ intraformational sandstone is an attractive secondary
API crude), on a down-plunge culmination on the objective in the Ucayali basin.
northward-plunging anticline that separates the east and Cretaceous sandstones sealed by overlying Creta-
west Ucayali basins (Figure 3). ceous or Tertiary shales are present in each of the sub-
Like the Marañon basin, most of the petroleum Andean basins. They are thin in the Madre de Dios basin,
discovered in the Ucayali basin is reservoired in Creta- but thick and more continuous in the Ucayali and
ceous clastics. There are, however, some important Marañon basins (Figures 9B, 10). The Aptian transgres-
exceptions. The Mipaya and Cashiriari gas fields have sive Cushabatay sandstones have a variable thickness
434 Mathalone and Montoya

SW YANAYACU 27X (PROJ.)


NE CHAMBIRA 123X (PROJ.)
SW NE
(SEC.) (SEC.)
0.0 0.5

1.0

1.0

2.0
2.0
Near Top Cretaceous
Near Top Cretaceous
Base Chonta Fm
Base Cretaceous 3.0
Base Chonta Fm

Base Cretaceous 3.0

4.0

0 1 2 km
0 1 2 km

4.0

AGUAYTIA 1X CORRIENTES 1X
W E SW NE
(SEC.) (SEC.)
0.0 1.0
0 1 2km

ceous
p Creta
Near To 1.0 1.5

ta Fm
Chon
Base

taceous
Base Cre 2.0 Near Top Cretaceous 2.0

ic Salt
Triass

Base Chonta Fm
3.0 2.5

Base Cretaceous

0 1 2 km

3.0

PARIAMANU 47-23-1X
SW NE
TAMBO SUR 4X (PROJ.) (SEC.)
SW NE
(SEC.)

Near Top Cretaceous 2.0 1.0

Base Chonta Fm Near Top Cretaceous

Base Cretaceous 3.0 2.0


Top Permian

Top Devonian

4.0 3.0

0 1 2 km 0 1 2 km

Figure 12—Seismic expression of the main structural styles of proven hydrocarbon traps. See Figure 16 for locations.
Petroleum Geology of the Sub-Andean Basins of Peru 435

HUASAGA 1X
W CAPAHUARI S-1X (PROJ.)
E
(SEC.) SW NE
(SEC.)
0.0
0.0

1.0

1.0
Near Top Cretaceous
Near Top Cretaceous 2.0

Base Chonta Fm

Base Cretaceous 3.0


Base Chonta Fm
Mid-Triassic 2.0

Base Cretaceous
4.0
m ent
Base
Top

0 1 2 km
0 1 2 km
5.0

3.0
MIPAYA
SW NE (SEC.)
AGUA CALIENTE 1X (PROJ.)
0.0
W E
(SEC.)
0.0
s
eou
etac
To p Cr
1.0 Near
ta Fm
Chon
Base 1.0
us
ta ceo
e Cre
Bas
Near Top Cretaceous
2.0 an
oni 2.0
e Dev
Bas
Base Chonta Fm
Base Cretaceous

3.0 3.0

0 1 2 km Base Devonian 0 1 2 km

Figure 12 (continued)

distribution but exceed 500 m in outcrop in the Cusha- porosities up to 25%. Permeabilities are variable, ranging
batay hills. Porosity generally varies between 10 and 22% from negligible to 1000 md. The Agua Caliente sand-
and permeability is moderate to good. The Cushabatay stones are sealed by shales of the lower Chonta
Formation has tested gas at 18 MMCFGD from the Formation. However, intraformational sandstones of the
Cashiriari-1X well in the southern Ucayali basin and oil Chonta Formation locally vary up to 300 m in thickness
at 2000 BOPD, where horizontal permeability is higher and form important reservoirs in the Marañon and
than 700 millidarcys (md) in the Agua Caliente field Ucayali basins. These shallow marine sandstones are
(Touzett, 1975). The equivalent Hollin Formation in the sealed by overlying intraformational shales that are also
Ecuadorian Oriente basin contains major oil reserves. capable of sealing large gas columns. Most of the
The Cushabatay shales of the overlying Raya Formation, reserves in the Corrientes field are reservoired in the
although locally interbedded with siltstones and sand- Chonta Formation (Cetico and Pona members), where
stones, are effective seals. For example, although the the porosities are about 21–23%. Horizontal permeability
Cushabatay is relatively sand-prone in the southern in this field (often over 1000 md) is controlled by deposi-
Ucayali basin, it is still capable of sealing large gas tional environment, such as marine barrier bars and
columns in highly deformed structures such as the fluvial channels. Reservoir continuity and connectedness
Cashiriari field. is good in the fluvial sandstones, but surprisingly poor in
The Agua Caliente Formation is commonly thickly the shorezone deposits. Figure 11 shows the stratigraphic
bedded, texturally and mineralogically mature, and has variation between the Forestal-1 and Corrientes 12 wells.
436 Mathalone and Montoya

The Campanian Vivian Formation is the most wide- of the Ambo Formation are best preserved in the basins
spread reservoir interval and varies between 20 and 160 of southern Peru (Figure 8A). Several organic-rich and
m in thickness (Figure 11B). The Vivian is attributed to coaly intervals with over 5% carbon are reported from
sedimentation in shorezone and fluvial depositional the Madre de Dios and South Ucayali basins where the
environments. It contains more than half of the oil in the Ambo Formation is a fairly ubiquitous gas and light oil
Marañon and Ucayali basins, mainly because of its strati- source. Like the Devonian source rocks, its distribution is
graphic position as the highest sealed reservoir in the patchy and maturity is high, usually over 2% Ro (Soto
Cretaceous succession. It is erosionally truncated over and Vargas, 1985), in the North Ucayali and southern
most of the Ecuadorian Oriente basin (where it is termed Marañon basins. The Upper Carboniferous and
the M-1 sandstone), and in the northeastern part of the lowermost Permian stratigraphy generally has lower
Marañon basin. This sandstone is generally homoge- source rock potential. The Copacabana Formation
neous, forming a continuous aquifer over most of the (Figure 5) consists mainly of lean platform carbonates,
basinal areas, although significant isolated units have with local interbedding of dark algal-rich shales in which
been described in the Shiviyacu field (Augusto et al., the TOC content in the overthrust belt of the western
1990). Reservoir parameters are generally excellent, and Madre de Dios basin may exceed 4 wt. %. The Copaca-
permeabilities are commonly over 1000 md; even sand- bana may therefore be a possible source of oil in the
stones with porosities less than 10% have acceptable South Ucayali and Madre de Dios basins.
permeabilities. The Vivian sandstone, sealed by Creta- The Copacabana Formation is overlain by hypersaline
ceous or Tertiary shales, is thus an excellent reservoir organic-rich marine shales of the middle Permian Ene
interval where strong water drive typically results in Formation. These shales are up to 300 m thick in outcrop
recoveries well over 40% of the original oil in place. and have an average TOC of 2–3 wt. % (maximum of
Clastic reservoirs occur throughout the Cretaceous 7 wt. %). Kerogen types I and II predominate, indicating
succession. Even below 4000 m depth, reservoir quality is a high oil-generating potential. The Ene Formation, like
favorable; the 1 md equivalent porosity cut off for the the other pre-Triassic units, has a patchy distribution
Vivian Formation is deeper than this depth. Very porous (Figure 7), but it is one of the most important oil source
Maastrichtian sandstones produce from the Maquia field rocks in the foreland basins of Peru.
in the Ucayali basin, and basal Tertiary and Paleozoic Overlying the prominent Middle Triassic unconfor-
sandstones are believed to have reservoir potential in the mity is the widespread Upper Triassic–Lower Jurassic
Marañon basin. Generally, however, Cretaceous clastics Pucara Formation, a westward-thickening wedge in the
are the primary focus for exploration in the Marañon and Marañon and Ucayali basins (Figure 8B). This formation
Ucayali basins, with upper Paleozoic rocks of secondary is absent from the eastern margin of the foreland basin,
importance. In contrast, the Paleozoic section is more including the South Ucayali and Madre de Dios basins.
important than the Cretaceous in the Madre de Dios The Pucara Formation comprises platform limestones,
basin due to better development of reservoir sandstones organic-rich limestones, and some interbedded organic-
in proximity to mature oil source rocks. rich shales. Individual units of organic-rich shales and
shaley limestones thicker than 50 m are exposed in the
Source Rocks Cushabatay Mountains, where the formation exceeds
1000 m in thickness (Figure 8B). Data are limited, but in
Potential oil and gas source rocks occur in the some outcrops, TOC content in samples ranges up to
Paleozoic and Mesozoic succession (Figure 5). Organic- 5 wt. % and samples are dominated by a sapropelic oil-
rich shale deposits occur in the Ordovician Contaya prone kerogen. It is believed that the Pucara Formation is
Formation and in Silurian deposits. These units, an important potential oil source rock in the western
however, are always overmature and unlikely to Marañon and Ucayali basins.
contribute substantially to the hydrocarbon system ( Soto Upper Cretaceous shales of the Raya and Chonta
and Vargas, 1985). More promising are the Devonian formations (equivalent to the Napo Formation of
shales of the Cabanillas Formation, which are thin and Ecuador) are part of the oil-prone shale blanket that
preserved as small remnants in the Marañon and North spans northwestern South America from Venezuela to
Ucayali basins (Figure 8A). In the South Ucayali and Peru. The principal oil-prone source rocks include the
Madre de Dios basins, the Devonian section can exceed Chonta Formation of the north Marañon basin and the
1500 m in thickness, with abundant but surprisingly Napo Formation of the Oriente basin, where the average
variable distribution of gas and liquid sources. (The TOC exceeds 3 wt. % (Figure 11A). There is also an
Devonian Los Monos and related shales are believed to isolated but important source area in the northwestern
be the source of the gas and liquids in the Bolivian Chaco Marañon basin (Figure 11A). These formations were
basin.) The Cabanillas shales in the overthrust belt have a deposited in a shelf setting that shoals southward in the
total organic carbon (TOC) content that locally exceeds Marañon basin. Their counterparts in the South Ucayali
3 wt. % and consists of structured and unstructured type and Madre de Dios basins are coarse-grained terrestrial
II kerogens. Some sections, however, have little source deposits with little source rock potential. Minor source
potential; the distribution of these low potential sections potential may exist in the shales of the Tertiary Pozo
is not clearly understood, but might be related to Formation, in which TOC content locally exceeds only
intraformational erosion. 0.5 wt. %. A few oil shows have been recorded from this
The Lower Carboniferous deltaic and marine clastics unit.
Petroleum Geology of the Sub-Andean Basins of Peru 437

Brazilian Brazilian
Shield 0.7 Shield
Maranon/Oriente Maranon
Basin Basin

COLOMBIA COLOMBIA

0.6
ECUADOR
ECUADOR
4.0
3.5
BRAZIL 1.1 BRAZIL
Iquitos Iquitos
4° PERU 4°
Ro (%)
3.

? IMMATURE
0

0. 0.6
6
1.0
OIL
0.8
2.
5

WINDOW
C.I. 0.5 °C/100m GAS
1.0
to a Nominal 0.8
3.5 20°C Surface C.I. = 0.1%Vitrinite Reflectance
Intercept
5
1. .0
1

0.8
Huallaga

0 .6
8° Huallaga
Pucallpa
Pucallpa 8°
Basin
Basin Ucayali Basin Ucayali Basin
Andes
3.
0

Mountains

0.6
Madre Madre
De Dios De Dios
Basin Andes Basin

0.8 1
Mountains
0.6

.0
BOLIVIA BOLIVIA

12° A Lima Ene


Basin 2.5 B Lima Ene
Basin
12°
2.

Pacific Pacific
0

Ocean Ocean
0 100 200 0 100 200

km km

75° 71° 75° 71°

Figure 13—(A) Geothermal gradient. (B) Geochemical maturity at the lowest Cretaceous interval, expressed as vitrinite
reflectance (% Ro).

Hydrocarbon Generation maturity of Cretaceous rocks exceeds vitrinite reflectance


values of 1.2% Ro, well above oil-generating maturities,
Geochemical analysis indicates that the depth to the over a significant part of the basin (Touzett and Sanz,
top of the oil-generative zone in the Marañon and Ucayali 1985).
basins varies from 2000 to 3300 m. This variation is attrib- Cretaceous source rock shales have the quality,
uted partially to the pattern of connate water convection maturity, and distribution to generate most of the
from the depocenter in the west. The geothermal gradient Marañon and Oriente oils. A large amount of geochem-
has a similar trend, with gradients as low as 2˚C/100 m ical work and oil to source rock correlation have been
adjacent to the deformation front and increasing eastward undertaken. Figure 14 shows that oils in the northern
up the foreland ramp to 3.5˚–4.0˚C/100 m. Most of the part of the Marañon basin and the Oriente basin correlate
measurements of bottom hole temperature are from with the Cretaceous Chonta and Napo shales, respec-
within or just below the Cretaceous succession, tively. Another large family of crudes from the southern
suggesting that these characteristics are partially due to part of the Marañon basin and the Ucayali basin were
convection within the Lower Cretaceous section. probably derived from evaporitic shales in the Permian
The map of present temperature at the base of the Ene Formation. Figure 11A shows the locations of the
Cretaceous succession (Figure 13A) indicates a large area Cretaceous-derived crudes, while Figure 7 illustrates the
of the Marañon basin that is above 140˚C. Smaller areas distribution of the Permian Ene Formation and the
in the Ucayali and Madre de Dios basins are also at this location of the crude oils it is believed to have generated.
temperature. The distribution of geochemical maturity These include substantial fields such as the 150-million-
(vitrinite reflectance) (Figure 13B) supports this trend, bbl Corrientes field, indicating significant oil generation
indicating a large “kitchen” area of mature Cretaceous from this formation. A significant potential source for
sedimentary rocks near the Ecuadorian border in north- which little data are available is the Triassic–Lower
western Marañon basin. Significant maturation of Creta- Jurassic Pucara Formation. A tenuous Pucara outcrop to
ceous deposits is attained only in small areas in the Ucayali oil correlation can be made (Figure 8B), and
Ucayali basin and not at all in the Madre de Dios basin. seeps along the northeastern margin of the Huallaga
Part of the Huallaga basin (Figure 13B) appears to have basin suggest a pre-Cretaceous source. Finally, seeps and
been deeply buried and recently uplifted, so that the oil shows in the Madre de Dios basin occur in Carbonif-
438 Mathalone and Montoya

STABLE CARBON ISOTOPE PLOT Brazilian


32 GROUP I Shield
Maranon/Oriente
PERMIAN Basin AREA OF MATURE OR
31 ENE FORMATION OVERMATURE SOURCE
SOURCE ROCKS ROCKS
TERRESTRIAL COLOMBIA Cretaceous
30 (Chonta/Raya/Napo)
Jurassic/Trias
(Pucara)
29 Permian
CRETACEOUS GROUP II (Ene)
C13 ALKANES

CHONTA (NAPO)
28 SOURCE ROCKS Iquitos Carboniferous
% PDB

DOMINANTLY 4° GROUP IIIA & Devonian


PERMIAN/JURASSIC? Ambo-Cabanillas
SOURCED OILS
27
Oil Group Source
I Cretaceous
26
II Permian
DOMINANTLY IIIA Jr./Tr./?Perm.
25 CRETACEOUS
SOURCED OILS IIIB Jr./Tr./?Perm.

BRAZIL
MARINE IV Perm./Carb./Dev.
Santiago
24 Basin
? V Carb./Dev.
23 24 25 26 27 28 29 30 8° Pucallpa
Aguaytia ?Perm./?Carb.
C13 AROMATICS % PDB Huallaga Condensate /?Dev.
Basin
Andes Ucayali Basin
Mountains
GROUP IIIB
UCAYALI BASIN MARANON BASIN CHONTA Madre
(TYPE III) JURASSIC TYPE I (CRETACEOUS De Dios
& PERMIAN SOURCE SOURCE) NW AREA GROUP IV Basin
MARANON BASIN VIVIAN TYPE I MARANON BASIN
(CRETACEOUS SOURCE) NW AREA TYPE II (PERMIAN
(ECUADOR BORDER) GROUP V BOLIVIA
SOURCE)
ECUADORIAN ORIENTE OILS SANTIAGO BASIN
OIL SEEPS (TYPE I) Lima
12° Ene
Basin
Pacific
Ocean
Figure 14—Oil to source correlation based on C13 stable 0 100 200

carbon isotope relationships. Other techniques have also km

been used in the Peruvian basins. (Ecuadorian data from 75° 71°
Dashwood and Abbotts, 1990.)
Figure 15—Oil generation, migration, and entrapment
domains.
erous and Devonian clastics and are related to intrafor-
mational source rocks.
Five groups of reservoired oils are recognized in sub- in the Huayuri field (Flores, 1991) supports this interpre-
Andean Peru (Figure 15). In the Marañon basin, large tation of remigration. Several authors (Sofer et al., 1985;
areas of mature to overmature Permian Ene Formation, Dashwood and Abbotts, 1990) suggest that much of the
Triassic–Jurassic Pucara Formation, and Cretaceous oil discovered in the Marañon and Oriente basins was
shales occur in the western depocenters. Ene strata also generated from the western margins of the basins before
fill rift basins beneath the eastern parts of the foreland Miocene deformation of the Eastern Cordillera inter-
basin. All the oils of the Ecuadorian Oriente basin and vened.
the northernmost Marañon basin are believed to be The oils of the Marañon and Oriente basins have a
related to a single Cretaceous group of source rocks broad range of API gravities (Figure 15), especially in the
(group I). A second group of oils in the southern Upper Cretaceous Vivian Formation. Only one of the
Marañon basin has strong Permian source rock affinities Chonta Formation oils (San Jacinto) is slightly biode-
(group II). Oils in the northern Ucayali basin show graded, but many Vivian Formation oils are severely
Permian and possible Triassic Pucara Formation influ- altered by bacterial degradation. The base of biode-
ences (groups IIIa and b). The large gas-condensate graded oil is at about 2800 m, significantly deeper than
reserves of the Cashiriari and San Martin fields appear to the 85ºC isotherm that is believed to exclude bacterial
be related to Permian and Carboniferous sources, with activity. It is inferred that significant migration and
some liquids possibly derived from the Permian Ene biodegradation predated the middle Miocene phase of
Formation (group IV). Finally, oils in the Madre de Dios basin subsidence.
basin appear to be derived from Devonian and Carbonif- A family of very light, shallow oils in the Oriente and
erous shales. Figure 15 summarizes these groups as well Ucayali basins occur in young thrust anticlines. Sofer et
as the distribution of the mature source rock intervals. al. (1985) convincingly argue for two phases of oil
The Peruvian foreland basin has a multiple source migration into the Vivian reservoirs of the northeastern
petroleum system that varies from basin to basin as well part of the Marañon basin. The salinity of formation
as within individual basins. Evidence of variation in water is generally a guide to meteoric water penetration,
crudes within the Cretaceous succession (group I) although not to its timing. There is a loose relationship
indicates separate pulses of primary migration and between decreasing API gravity due to biodegradation
possibly remigration. A tilted residual oil–water contact and decreasing salinity of formation water.
Petroleum Geology of the Sub-Andean Basins of Peru 439

Analysis of potentiometric gradients shows a weak Brazilian


Shield
basinward flow from the Brazilian shield. For the Vivian
Maranon
aquifer, a 150-m drop generally occurs in the potentio- Basin
metric surface from the edge of the Brazilian shield COLOMBIA
westward to the center of the Marañon basin. There is no ECUADOR
CAPAHUARI
evidence for basinward fluid flow eastward from the
BRAZIL
Andean foothills, with the exception of strong meteoric CORRIENTES
water influx northward from the Cushabatay and Iquitos Seismic Example
Contaya mountains that border the basin on the south. TAMBO Dominant Trap Type Fig.12

Few data are available from the Ucayali basin, but it HUASAGA CHAMBIRA Inversion Traps
appears that Cretaceous, and in particular Vivian Drape/Compression Traps
Some Salt Effects
Formation, waters are mostly fresh. Water associated YANAYACU
Basement Involved
with the Maquia, Agua Caliente, and Aguaytia fields Thrust Traps
have NaCl concentrations less than 1000 ppm (Ciquero, Detached
1982). In the Ucayali basin, little biodegredation of Thrust Faulted/Folded
Traps
hydrocarbon liquids has occurred, although aquifers Structural
adjacent to these accumulations contain meteoric water 8°
Trend
Pucallpa
at most depths (e.g., Maquia 37º API). This contrasts with AGUAYTIA
Ucayali Basin
the Marañon basin, where there have been several
episodes of biodegradation of reservoired oils. Andes AGUA CALIENTE
Mountains
Madre
de Dios
Structural Trap Styles Basin
MIPAYA
BOLIVIA
Well over 1 billion bbl of crude oil and condensate
have been discovered in the composite foreland basin of 12° Lima Ene PARIAMANU
Basin
Peru. Hydrocarbons have been trapped in a variety of Pacific
structures, many of which are young and were created or Ocean
substantially modified by late Miocene–Pliocene 0 100 200

Quechua III tectonism. km

75° 71°
The Marañon basin is a westward-facing composite
foreland basin that was built on earlier extensional and Figure 16—Structural provinces, showing distribution of
transtensional basins. The late Tertiary–Pleistocene fold trap types. Location of seismic field sections of Figure 12
are indicated.
and thrust belt forms the foothills and isolates the Ene,
Huallaga, and Santiago basins (Figure 3). The dominant
structural style of the eastern ramp of the foreland basin has had a more complex history, possibly involving
consists of reactivated and inverted normal faults of Eocene Incaic compression before Miocene Quechua III
Paleozoic age (Figure 16). Toward the west, these older modification. The Aguaytia gas field in the Ucayali basin
extensional structures were depressed directly by the illustrates the modifying effects of Triassic salt flow on an
applied overthrust load without inversion (Figure 12). otherwise simple compressional structure. Structures
The sub-Andean foothills can also be divided into two such as South Tambo are subtle drape folds formed
structural domains. To the west, faulting is dominated by above older structures. Many of these low-amplitude
thrusts that detach in Triassic salt in the north or folds are characterized by slight thinning of the Creta-
Devonian shale in the south. Toward the east, an uplifted ceous section over their crests, suggesting the onset of
zone is dominated by basement-involved thrusting compression in the middle Cretaceous. The large 150-
(Figure 16). million-bbl Capahuari field is an example of this struc-
The Ucayali basin is generally more deformed than tural type.
the Marañon basin. Detached thrusts continue Late extensional movement and differential
southward from the Marañon basin, but the Ucayali compaction of the Paleozoic section in the western
basin is surrounded on all sides by basement-involved foreland created forced folds in the Cretaceous and
thrusting (Figures 3, 16). The sub-Andean thrust belt Tertiary sections. This has resulted in low-amplitude
continues southward from the Huallaga basin to the closures where forced folds are superimposed on a
Camisea area at the southern end of the Ucayali basin regional dip. The Huasaga structure is an example
where it is separated from the Ene basin by the reacti- (Figure 12). Withjack (1989) has described the dynamics
vated Shira high. In contrast, the Madre de Dios basin of forced folding. The foreland of the Madre de Dios
comprises a largely undeformed southward-dipping basin is virtually unstructured as illustrated by the
foreland ramp in the north and a thin belt of foothills seismic line through the Pariamanu structure in Figure
thrust structures in the south. 12. The thin Cretaceous section is little deformed except
Figure 12 shows the seismic expression of these struc- for forced folds.
tural styles. The Yanayacu and Chambira fields indicate Thrust deformation includes basement-involved
the youthfulness of inversion, which involved much of structures such as the Agua Caliente field in which oil is
the Tertiary section. The 150-million-bbl Corrientes field trapped in the basal Cretaceous Cushabatay Formation,
440 Mathalone and Montoya

and thin-skinned thrusting, which created a series of of thrust belt loading (Quechua III tectonism). The result
closed anticlines such as the Camisea and Mipaya fields of this deformation has been structural trap formation
(Figure 12). Figure 16 shows the regions of Peru where and rapid burial maturation imposed by the thick
each of the four main structural types predominate. Tertiary molasse wedge. Due to the relative youth of
some of the petroleum systems, together with remigra-
tion of oils generated earlier, very young structures are
capable of trapping giant hydrocarbon accumulations.
CONCLUSIONS The sub-Andean foreland basins of Peru cover an area
Slightly over 5 billion bbl of recoverable hydrocarbon of more than 370,000 km2. With a density of only one
liquids have been discovered in the Ucayali-Marañon wildcat well per 3000 km2, a substantial amount of
basin complex, including the contiguous Oriente and exploration is required to realize Peru’s petroleum
Putumayo basins of Ecuador and Colombia, respectively. potential.
Over 1 billion bbl of these reserves have been found in
the Peruvian Marañon and Ucayali foreland basins by
113 wildcat wells. The success rate is about 25%. Few
wells have been drilled, and no substantial discoveries Acknowledgments Most of the information presented in this
have been made on the Peruvian side of the Madre de paper was provided by Petroperu, to whom the authors extend
Dios basin. their thanks. Permission to publish was also granted by
Hydrocarbon source rocks in the sub-Andean basins Eurocan Ventures, for whom some of the work had been done
of Peru range from Devonian to Late Cretaceous in age. on a proprietary basis. Thanks are extended to Antoine Fabre,
Most of the oil and gas discovered is in Cretaceous also of Eurocan Ventures, for his assistance. The authors would
terrigenous clastic reservoirs, although there are also also like to thank the reviewers, George Kronman, Robert
significant Paleozoic prospects. The petroleum is Meneley, and particularly Tony Tankard, for their constructive
contained in four principal trap types. Basement- help. Finally, the authors would like to express their gratitude
involved or thin-skinned thrusts dominate the western to Jo-Lynn Cole who created all the artwork in this paper.
basin tracts and the Andean foothills, respectively,
whereas the eastern foreland is characterized by reacti-
vated older faults and the central area by forced folds
and subtle anticlines. REFERENCES CITED
In most of these structural provinces, trap formation
and modification is relatively young, being related Augusto, M., C. Ardiles, and C. Orosco, 1990, Geológica del
mainly to the late Miocene Quechua III tectonic episode. yacimiento Shiviyacu: Boletin de la Sociedad Geológica del
Peru, v. 81, p. 63–80.
An exception is the central-western Marañon basin of Canfield, R. W., 1991, Sacha field-Ecuador: Oriente basin, in
Peru and the central-western part of the Ecuadorian N. H. Foster and E. A. Beaumond, eds., Atlas of oil and gas
Oriente basin. This region is characterized by older fields, structural traps V: American Association of
compressional anticlines and by forced folds that Petroleum Geologists, Treatise of Petroleum Geology,
generally predate the late Oligocene and may be as old p. 285–305.
as middle Cretaceous in places. Many of the large Ciquero, J. Z., 1982, Future of petroleum exploration in the
Ecuadorian fields such as Sacha ( 750 million bbl recover- sub-Andean basins of Peru: Petroleum exploration in the
able) (Canfield, 1991) are of this structural type. sub-Andean basins of Venezuela, Colombia, Ecuador and
The other three structural provinces, although Peru, Symposium, Bogota, Colombia.
comprising very young structures, each have examples Dalmayrac, B., G. Lambacher, R. Marocco, C. Martinez, and B.
Tomasi, 1980, La chaine hercynienne d’America du Sud-
of large hydrocarbon accumulations. The Cashiriari and structure et evolution d’ orogene intracratonique:
San Martin gas fields of the southern Ucayali basin have Geologigische Rundschan, v. 69, p. 1–21.
proven reserves of about 7 tcf of gas and 400 million bbl Dashwood, M. F., and I. L. Abbotts, 1990, Aspects of the
of condensate trapped in young compressional anticlines petroleum geology of the Oriente basin, Ecuador, in J.
associated with detached thrusts. So far in Peru, only Brooks, ed., Classic petroleum provinces: Geological
small accumulations such as Maquia and Agua Caliente Society of London, Special Publication 50, p. 89–117.
have been found in anticlines associated with basement- Eduardo, H., 1991, Paleogeogrifía del Paleozoico en el Oriente
involved deformation. Farther north, however, anticlines Peruano: VII Congreso Peruano de Geología, Symposium,
such as this trap giant oil accumulations in Ecuador and Lima, p. 269–276.
the Putumayo basin of Colombia (e.g., the Orito field Flores, A. B., 1991, Campos Huayuri (norte y sur)–Bloque 1-
AB. Un ejemplo de basculamiento del area in el Terciario
with over 200 million bbl of oil). The eastern foreland of Superior migración secundaria de petróleo: VII Congreso
the Marañon basin contains the Corrientes field with Peruano de Geologica, Lima, Peru, p. 239–244.
some 150 million bbl of recoverable oil trapped in a rela- Martinez, C., 1980, Structure et evolution de la chaine
tively young anticline associated with an older fault that Hercynienne et de la chaine Andine dans le nord de la
has been reactivated in a reverse sense (Figure 12). Cordillere des Andes de Bolivie: Oravour et Documents de
Typical of the sub-Andean tract of foreland basins, the l’Orstom, Paris, v. 119, 352 p.
foreland basins of Peru are comparatively youthful and Mégard, F., 1978, Le Substratum Paléozoique, La Chaine
are attributed to the middle Miocene and Pliocene phase Hercynienne, in Etude Geologique des Andes du Pérou
Petroleum Geology of the Sub-Andean Basins of Peru 441

Central: Contribution á l’Etude Geologique des Andes, No. Soto, F., and J. Vargas, 1985, Posibilidades hidrocarburíferas
1 Memoires Orstom, no. 86. del Precretacico en el oriente del Peru: Boletin Tecnico
Mégard, F., 1984, The Andean orogenic period and its major ARPEL, v. 14, p. 131–143.
structures in central and northern Peru: Journal of the Touzett, P., 1975, Evaluacion geológica del yacimiento petro-
Geological Society of London, v. 141, p. 893–900. lifero Agua Caliente: Boletin de la Sociedad Geológica del
Mendez, J., A. Navarini, D. Plaza, and V. Viera, 1972, Faja Peru, v. 48, p. 9–24.
eruptiva de la Puna oriental: Actas Congreso Geologico Touzett, P. J., and V. R. Sanz, 1985, Presente y futuro de la
Argentino, Buenos Aires V, v. 4, p. 83–100. exploración petrolera en las cuencas sub-Andinas del Peru:
Mukasa, S. B., and D. J. Henry, 1990, The San Nicolas Exploración Petrolera en las Cuencas Sub-Andinas, II
batholith of coastal Peru: early Paleozoic continental arc or Simposio Bolivariano, Bogota, Colombia, 70 p.
continental rift magmatism?: Journal of the Geological Verner, R., and J. P. Xavier, 1990, Using remote sensing data
Society of London, v. 147, p. 27–39. for hydrocarbon exploration in the Andean basins: Bulletin
Parra, V. S., 1974, Geologica preliminar del area Tigre-Corri- de Centre Recherche Elf Aquitaine, v. 14, p. 403–418.
entes: Boletin de la Sociedad Geológica del Peru, v. 44, Withjack, M. O., K. E. Meisling, and L. R. Russell, 1989, Forced
p. 106–127. folding and basement-detached normal faulting in the
Petroperu, 1989, Peruvian Petroleum, a renewed exploration Haltenbarken area, offshore Norway, in A. J. Tankard and
opportunity: Petroperu, Lima, 101 p. H. R. Balkwill, eds., Extensional tectonics and stratigraphy
Portugal, J. A., and L. Gordon, 1972, Geological history of of the North Atlantic margins: AAPG Memoir 46,
southern Peru: II Congresso Latino Americano de p. 567–575.
Geològica, Cordillera Andina, 42 p.
Rodriguez, A., 1982, Petroleum exploration in the Santiago
basin: Petroleum exploration in the sub-Andean basins of
Venezuela, Colombia, Ecuador and Peru, Symposium,
Bogota, Colombia, 10 p.
Authors’ Mailing Addresses
Salas, G., 1991, Factores geologicos de control de acumula- Jeremy M. P. Mathalone
ciones de Hidrocarburos en las cuencas del Oriente PetroSantander Inc.
Peruano: Exploracion petrolera en las cuencas subandinas, 5847 San Felipe, Suite 1650
Simposio Bolivariano IV, Bogota, Colombia, trabajo 29. Houston, Texas 77057
Sempere, T., 1995, Phanerozoic evolution of Bolivia and U.S.A.
adjacent regions, in A. J. Tankard, R. Suarez, and H. J.
Welsink, Petroleum basins of South America: AAPG
Memoir 62, this volume. Manuel Montoya R.
Sofer, Z., J. E., Zumberge, and V. Lax, 1985, Stable carbon CIA Consultora de Petroleo S.A.
isotopes and biomarkers as tools to understanding genetic Mariano de los Santos 198
relationship, maturation, biodegradation and migration of Of. 405, San Isidro
crude oils in the northern Peruvian Oriente (Marañon) Lima 27
basin: Organic Geochemistry, v. 10, p. 377–389. Peru
442 Mathalone and Montoya

Appendix
Figure A.1 shows the distribution of exploration wells in the (A) Marañon, (B) Ucayali, and (C)
Madre de Dios basins. This information is provided as a basic reference. Tables A.1 and A.2
summarize the oil and gas fields discovered in the Marañon and Ucayali basins, respectively,
including the producing stratigraphy, typical porosities, and API gravities.

Colombia
Sta. Lucia
Ecuador MARANON

seo
PANTOJA BASIN

Abi
MARANON

Rio
Sta. Clara-1 Rayo-1
BASIN 7°
Cunambo-3X
2°S Huaya-3X
San Cachiyacu-1
Jacinto Huaya-4X Maquia
Tangarana-4X Ponasillo
Forestal Ext 1X Inuya-1
SANTIAGO Carmen Amaquiria-32X Brazil
Shiviyacu Forestal HUALLAGA
BASIN Macusari Pilar-1 Bartra Pacaya
Cashiboya-1
Huayuri Jib Ext Tigre-22X BASIN Cashiboya-29X
Capahuari N. Jibaro Pisque-1 Tiruntan-1X
Jibarito
Capahuari Sur. Margarita Coninca-2 8°
Andoas-52X Martha
Maynas Tahuaya-1
Tambo Bolognesi-62X
Yanez-14X Sur Dorisa Plantayacu-83X Aguaytia-1

Rio
Huasaga-1X Ceci Valencia Nva. Esperanza
Tunchiplaya-95X
Otorongo Pucacuro-1 Aguaytia-3X Zorrillos PUCALLPA

Hu
Chapuli-1X Pavayacu Huangana Nanay-26X

all
Huitoyacu Capirona S. Juan Neshuya 5-1

ag
Intuto-23X Belen-4X Tamaya-2X

Rio
Sungachi-2X Copal-15X IQUITOS
Chambira-4X Aerico-18X Agua Caliente Platanal-1X

Uca
Corrientes 9°
Piuntza Chambira mazon Chonta-1
Rio A 4°S Sanuya-3X

yali
Tigrillo-3X Este-123X
a
astaz

La Colpa-1X
Dominguza Patoyacu-36X Nahuapa-24X
Ungumayo-1X Rio Caco-4X
Rio P

Mahuaca-3X Nucuray Maranon 10-1


Cuinico
Norte-4X Concordia-17X Runuya-1X UCAYALI
li

Cuinico Sur-1X BASIN


ya

Maranon 8-1
ca

HUANUCO Oxapampa 7-1


oU

Pauyacu-1X Maranon 22-1 10°

Rio Uca
Ri

Yanayacu-32X Oxapampa 19-1 Oxapampa 13-2


Bretana-1X Yarina-2X Oxapampa 7-1
AN Pastococha-6X Oxapampa 7-2 MADRE DE

yali
Yanayacu
A

DE Envidia-4X DIOS BASIN


N

AN Viracocha-7X Samiria
D

OXAPAMPA
EA B

BE MO Zapote-3X
Samiria Sur-3X ENE
N EL

LT UN
M T

TA Yurimaguas-1X BASIN
O
U

IN 11°
N

Loreto-1X Tapiche-2X
TA

YURIMAGUAS Tamanco-1X Palmera-4X 6°S Sepa-1x


IN

MOLLOBAMBA Palo Seco-1X Sta. Elena

Rio
Shanusi-2X CAMISEA
La Frontera-3X Mipaya-5X

En
LA OROYA San Martin-1X

e
Sta. Lucia
Oil well
A
TARAPOTO Oil well
Dry well B Ri
oM
Segakiato-2X
Dry well Brazil
HUALLAGA Sta. Clara-1 Rayo-1 Oil shows an Armihuari-4X 12°
Oil shows tar
BASIN Huaya-3X Gas Condensate well o Cashiriari-3X
Gas Condensate well
Huaya-4X 13°
76°W 74°W 76°W 75°W 74°W 73°W
0 20 40 60 80 100 km
Refinery Trans Andean 0 20 40 60 80 100 km
Pipeline Refinery

MARANON BASIN UCAYALI BASIN

73°W 72°W 71°W 70°W 69°W 68°W 67°W

Ur FITZCARRALD Brazil
ub ARCH
am
ba Bolivia
11° R. Peru
UCAYALI
BASIN MADRE DE
CAMISEA DIOS BASIN Pando
San Martin-1X
M Manuripi
Mipaya Segakiato-2X an Rio Cariyacu
uR
ive
12° r Puerto Primo
Cashiriari-3X
Los Amigos
Pariamanu
Armihuari-4X Madre
De Dio Bolivia
s River
er iv

C
Beni R

13°
Karene FORELAND
EASTERN
CORDILLERA

SU
14° CUZCO B- MADIDI
AN HIGH
DE
AN
TH
Oil well RU
Dry well 0 20 40 60 80 100 km ST
BE
Oil shows LT
15°
Gas Condensate well
Tuichi

MADRE DE DIOS BASIN


Figure A.1—Distribution of exploration wells in the (A) Marañon, (B) Ucayali, and (C) Madre de Dios basins.
Petroleum Geology of the Sub-Andean Basins of Peru 443

Table A.1—Oil Fields of the Marañon Basin

Field Operator/ Main Porosity Gravity Estimated Original Oil


Year Discovered Reservoirs (%) (°API) Recovery (%) Reserves (MMbbl)
Bartra Occidental 1974 Vivian 21 11 11 ?
Chonta 14 20 10 ?
TOTAL 26.1
Bretana Amoco 1974 Vivian 22 13.7 25 23.0
Capahuari S. Occidental 1973 Vivian 16 38 50 141.4
Chonta 12 26 15 11.1
TOTAL 152.5
Capahuari N. Occidental 1972 Vivian 16 36 15 0.8
Capirona Petroperu 1972 Chonta 18 30 25 4.7
Carmen Petroperu 1974 Vivian 16 20 25 1.3
Ceci Petroperu 1981 Chonta 12 43 30 0.1
Chambira E Petroperu 1989 Chonta 17 27 25 15.4
Corrientes Petroperu 1971 Basal Tertiary 22 1.4
Vivian 21 24 30 1.5
Chonta 18 28 35 143.2
TOTAL 146.1
Dorissa Occidental 1978 Vivian 16 36 55 45.5
Chonta 15 42 49 6.2
TOTAL 51.7
Forestal Occidental 1973 Vivian 17 20 35 34.3
Chonta 17 38 40 12.7
TOTAL 47.0
Huayuri S. Occidental 1977 Vivian 16 33 24 13.9
Chonta 15 37 23 10.1
TOTAL 24.0
Jibarito Occidental 1981 Vivian 19 11 12 23.7
Jibaro Occidental 1974 Vivian 19 10 10 16.2
Pavayacu Petroperu 1972 Vivian 19 46 42 7.0
Chonta 17 29 26 16.4
TOTAL 23.4
Samiria S. Petroperu 1975 Vivian 22 20 30 19.0
San Jacinto Occidental 1978 Vivian 21 14 6 27.4
Chonta 20 32 26 22.4
TOTAL 49.8
Shiviyacu Occidental 1973 Vivian 17 19 29 81.0
Chonta 17 40 43 11.9
TOTAL 92.9
Tambo S. Occidental 1982 Cushabatay 12 45 25 4.2
Valencia Nva. Petroperu 1975/80 Vivian 22 46 35 4.6
Esperanza
Chonta 18 45 30 2.5
TOTAL 7.1
Yanayacu Petroperu 1974 Vivian 20 19 28 5.2
Chonta 16 19? 28 0.1
TOTAL 5.3
GRAND TOTAL 734.3
444 Mathalone and Montoya

Table A.2—Oil and Gas Fields of the Ucayali Basina

Field Operator/ Main Porosity Gravity Estimated Original Gas Original Oil
Year Discovered Reservoirs (%) (°API) Recovery (%) Reserves (tcf) Reserves (MMbbl)
Agua Ganso Azul Cushabatay 23 44 45 — 13.0
Caliente 1939
Raya 18 22 1.7
TOTAL 14.7
Aguaytia Mobil 1962 Cushabatay 18 68 55 0.25 20.0
Cashiriari Shell 1986 Vivian 18 64–66 65 (gas) 1.4 77.0
Chonta 19 64–66 65 (gas) 1.2 66.0
Agua Caliente 17 64–66 65 (gas) 1.9 105.0
Cushabatay/Ene 14 64–66 65 (gas) 0.3 17.0
TOTALS 4.8 265.0
Maquia Oriente 1957 Casablanca 30 37 33 5.0
Vivian 30 37 46 13.1
TOTAL 18.1
Mipaya Shell 1987 Ene 12 60 70 (gas) 0.1 4.0
Pacaya Petroperu 1982 Cachiyacu 16 31 35 0.4
San Martin Shell 1984 Agua Caliente 17 60–62 70 (gas) 1.0 88.0
Cushabatay 13 60–62 70 (gas) 0.6 53.0
TOTALS 1.6 141.0
GRAND TOTAL 6.75 463.2
aAll reserves quoted are proven. Probable and possible liquid reserves of the Cashiriari and San Martin Fields are in the order of the proven reserves.
Petroleum System of the Northern and Central
Bolivian Sub-Andean Zone

P. Baby R. Limachi
ORSTOM
Grenoble, France E. Mendez
YPFB
Santa Cruz, Bolivia
I. Moretti
IFP
Rueil Malmaison, France J. Oller
Petrolex
Santa Cruz, Bolivia
B. Guillier
ORSTOM
Quito, Ecuador M. Specht
Total
Paris-LaDefense, France

Abstract

A coupled study of the kinematics of thrusting and hydrocarbon maturation has been carried out in the
northern and central sub-Andean belt of Bolivia to define the petroleum potential of the area. In
addition to the classic Devonian source rock (Tomachi–Tequeje formations to the north and Iquiri–Limoncito
formations in the central area), two other source rock intervals are recognized: the Retama Formation (Upper
Devonian–Lower Carboniferous) and the Copacabana Formation (Upper Carboniferous–Lower Permian).
These are the most prospective units in northern Bolivia and are of marine origin. The structural style varies
from north to south due to variations in the sedimentary column involved in the thrusts. The orogenic front
was guided by the northern boundary of a Paleozoic sedimentary wedge. In the Boomerang area, this
boundary is oriented obliquely to the regional shortening and controlled the development of a prominent
transfer zone. To the north, the thrusts are wider and the amount of shortening increases. The western part of
the northern sub-Andean zone is characterized by a very thick Tertiary piggyback basin fill.
Two phases of hydrocarbon maturation are recognized. The first began in Early Carboniferous and affected
mostly Devonian strata. Formation of structural traps during this period occurred rarely. The entire basin was
then deeply eroded in Permian–Jurassic time, causing any hydrocarbons that may have formed to be lost. The
second phase of maturation was contemporaneous with Andean deformation and with the resulting burial
under the Tertiary cover in the foreland basin and in piggyback basins on thrust structures. The hydrocarbon
expelled during this period may fill the Andean anticlines. The known source rocks are not proven to be gas
prone, but current discoveries indicate a high gas to oil ratio that may be due to secondary cracking in the
source rock. Because the initial potential of the source rocks is low, expulsion of heavy compounds is expected
to be weak.

Resumen

U n estudio combinado de la cinemática de los corrimientos y de la maduración de los hidrocarburos ha


sido realizado en el sub-Andino norte y centro de Bolivia a fin de definir el potencial petrolífero de la
zona. Además de las clásicas rocas madres devónicas (formaciones Tomachi–Tequeje en el norte y formaciones
Iquiri–Limoncito en el centro), dos otras han sido definidas: la Formación Retama (Devónico sup. a
Carbonífero inf.) y la Formación Copacabana (Carbonífero superior a Pérmico inferior) que presentan el mejor
potencial en el sub-Andino norte. Son de origen marino. El estilo estructural cambia de Norte a Sur, debido a
las variaciones de espesor y de litología en la columna sedimentaria implicada en los corrimientos. El frente
orogenico fue guiado por el borde norte de una cuña sedimentaria paleozoica. En la zona del Boomerang, este
borde esta orientado oblicuamente en relación con la dirección regional de acortamiento y controló el desar-
rollo de una prominente zona de transferencia. Hacia el norte, los corrimientos son mas anchos y el acor-

Baby, P., I. Moretti, B. Guillier, R. Limachi, E. Mendez, J. Oller, and M. Specht, 1995, 445
Petroleum system of the northern and central Bolivian sub-Andean zone, in A. J.
Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG
Memoir 62, p. 445–458.
446 Baby et al.

tamiento aumenta. La parte oriental del sub-Andino norte se caracteriza por un espeso relleno sedimentario
terciario contemporáneo de la deformación andina.
Dos fases de maduración de hidrocarburos han sido reconocidas. La primera empieza al principio del
Carbonífero donde la columna sedimentaria paleozoica es espesa; afecta principalmente al Devónico. Las
trampas estructurales de este período son muy raras. Entre el Pérmico y el Jurásico, la cuenca ha sido profunda-
mente erosionada, y los hidrocarburos generados probablemente han desaparecido. La segunda fase de madu-
ración es contemporánea de la deformación andina. Se debe a la fuerte subsidencia y sedimentación terciaria en
la cuenca de ante-país y en las cuencas transportadas por los corrimientos. Los hidrocarburos expulsados
durante este segundo período pueden haber rellenado anticlinales andinos. No se ha mostrado que las rocas
madres conocidas son propensas a generación de gas, pero los hidrocarburos descubiertos indican una alta
proporción gas-petroleo. Se interpreta como el resultado de un crácking secundario en la roca madre. El bajo
potencial inicial de las rocas madres hace que la expulsión de los componentes pesados debe ser reducida.

INTRODUCTION
As in many compressional areas, the petroleum
potential of the sub-Andean zone of Bolivia depends on
the relative timing between structuring and maturation.
All the known source rocks are Paleozoic in age.
Therefore, a study of the petroleum system requires a
study of the geology, the kinematics of thrust emplace-
ment, and the kinetics of the various source rock matura-
tions. The results of these studies are presented. Differ-
ences in the sedimentary columns and structural styles
from north to south are emphasized, and the potential
source rocks are defined and their maturation docu-
mented. In some cases, we address the maximum
amount of burial before the Andean deformation; there is
considerable uncertainty because of the amount of
erosion.

REGIONAL SETTING
The sub-Andean zone of Bolivia is a complex foreland
fold and thrust belt (Roeder, 1988; Sheffels, 1988, 1990;
Baby et al., 1989, 1992, 1993) that forms the eastern edge
of the central Andes mountains (Figure 1). It is bounded
along the edge of the Cordillera Oriental by the Main
Frontal thrust (CFP), whereas the orogenic front extends
Figure 1—Simplified tectonic map of Bolivia showing the
below the Beni and Chaco plains to the east. Deformation
location of the sub-Andean zone, a complex fold and thrust
started in the late Oligocene and is continuing today belt, and the studied cross sections. In its central part, the
(Sempere et al., 1990). The material involved in sub- Bolivian sub-Andean zone forms a bend (Santa Cruz bend)
Andean thrusting in Bolivia consists of an Ordo- characterized by the prominent Boomerang-Chapare
vician–Cretaceous series and an upper Oligocene–Recent transfer zone. In the northern and central part, the propaga-
continental foredeep fill. The preorogenic sedimentary tion of the orogenic front was guided by the northern
series show lateral variations in facies and thickness that boundary of the Paleozoic sedimentary wedge. SC, Santa
play an important role in controlling the structural Cruz; CFP, Main Frontal thrust; BCTZ, Boomerang-Chapare
geometry (Baby et al., in press). In its central part transfer zone.
(16°–17° S lat), this fold and thrust belt forms a bend
(Santa Cruz bend) characterized by a prominent transfer southern sub-Andean zone (18°–22° S lat), which is
zone. From north to south, the structural geometry oriented north-south. This southern sub-Andean zone is
displays variations in the amount and direction of short- not discussed in this paper.
ening (Figures 1, 2).
Three structural zones are recognized: (1) the northern Northern Sub-Andean Zone
sub-Andean zone (13°–17° S lat), which is oriented
northwest-southeast; (2) the central sub-Andean zone The Paleozoic succession (Figure 3) is nearly complete
(17°–18° S lat), which changes from a northwest- for the Ordovician–Permian and is unconformably
southeast to a north-south orientation; and (3) the overlain by a Mesozoic sandstone cover up to 800 m
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 447

Figure 2—Cross sections constructed and balanced from field studies and subsurface data. The structural geometry shows
important variations from north to south. The preorogenic sedimentary succession shows lateral variations of facies and
thickness that are important in controlling the structural geometry. C.F.P., main frontal thrust. See Figure 1 for locations.

thick. Toward the northeast, the thickness of the Ordovi- of the northern sub-Andean zone is characterized by a
cian decreases, the Silurian series disappears, and the very thick (6500–7000 m) Tertiary piggyback basin fill
Devonian, Carboniferous, and Permian successions (Alto Beni syncline). The main décollement level is at the
progressively wedge out. At the top of the preorogenic base of the Paleozoic column (Ordovician shales). The
stratigraphic column, the continental foreland deposits other décollement levels are shallower and are located in
are up to 5000 m thick. Devonian, Carboniferous, and Permian shales. The
The thrusts are wide, with their wavelengths always foredeep basement slopes at 4°. The maximum amount
being more than 10 km (Figure 2A, B). The western part of shortening is 135 km.
448 Baby et al.

in press). The major décollement is located at the base of


the Paleozoic sedimentary wedge which slopes at 10°.
The maximum amount of shortening is 60 km.

SOURCE ROCKS
Due to facies variations, a large number of formation
names have been introduced into the literature. Never-
theless, the stratigraphic column has been recently
rearranged and published by YPFB-ORSTOM (Sempere,
1990; Oller, 1992). We use this stratigraphic framework to
date the source rock formations.
To describe the various source rocks fully, the
Yacimientos Petroliferos Fiscales Bolivianos (YPFB)
database has been used and augmented with new data
collected by sampling outcrops and wells. The classic
source rock of the Bolivian sub-Andean zone is the
Devonian and probably the Silurian, but two other
Paleozoic formations are believed to have some
potential. Locally, an Upper Cretaceous source rock
(Flora Formation) with limited potential is known in the
northern sub-Andean zone.

Silurian
The Silurian source rock, of Wenlockian–Prodolian
age (430–410 Ma), reflects siliciclastic deposition on a
marine platform. The Upper Silurian is known in the
southern sub-Andean zone and foreland, in the central
part, and in a small part of the northern sub-Andean
zone, as shown in Figure 5A. Its thickness increases from
north to south, where it is up to 1500 m thick. The few
samples available to us do not permit a detailed evalua-
tion or even a confirmation of the original petroleum
Figure 3—Generalized stratigraphic column for the potential of this interval. Nevertheless, marine type II
northern sub-Andean zone of Bolivia. Two main erosional source rock characteristics are used for the modeling
unconformities are present in the Triassic (about 205 Ma) procedure to quantify the maturity of a hypothetical
and upper Oligocene (27 Ma) sections. The latter resulted
from the beginning of Andean deformation. The Middle
Silurian source rock.
Triassic erosion probably occurred during an episode of
extension related to the initiation of Gondwana fragmenta- Devonian
tion. Two prominent nondepositional unconformities are
present in the Cretaceous (144–68 Ma) and Paleocene The Devonian succession is Lochkovian–Famenian in
(53–27 Ma) sections. age (410–360 Ma). Its depositional setting was essentially
the same as during the Late Silurian (marine platform
deposition). Its thickness increases toward the south,
Central Sub-Andean Zone where it is 1500 m thick south of the Boomerang. Its
regional distribution is shown in Figure 5B. The basin
The preorogenic stratigraphic column is characterized extends northward into Peru and southward into
by a Paleozoic sedimentary wedge (Figure 4). It is a Argentina and Paraguay. The Devonian has been eroded
continuous succession, ranging from Ordovician to to the northeast; the age of erosion is discussed later.
Carboniferous in age and thinning northward onto the Source rocks include the Tomachi and Tequeje forma-
Precambrian–Cambrian Brazilian shield. It is uncon- tions in the north and the Iquiri and Limoncito forma-
formably overlain by 500 m of Mesozoic rocks and more tions in the central part of the basin.
than 1600 m of upper Oligocene–Recent deposits. We have not found immature Devonian samples and,
The orogenic front (Figures 1, 2C) is characterized by consequently, have not determined the kinetic parame-
the Boomerang-Chapare transfer zone, which is inter- ters. Since the depositional environments of the Tomachi
preted as an oblique ramp. The propagation of the and Iquiri formations are the same as the Retama
orogenic front was guided by the northern boundary of Formation of Carboniferous age, the type Retama section
the Paleozoic sedimentary wedge, which is oriented described below is used to model the Upper Devonian
obliquely to the regional shortening direction (Baby et al., source rocks. The Tequeje and Limoncito formations
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 449

Figure 4—Sedimentary wedge of the Boomerang area, in the central sub-Andean zone. The two main erosional unconformi-
ties (Triassic and upper Oligocene) and the two prominent nondepositional unconformities (Cretaceous and Paleocene) are
also present in this region. See Figure 1 for location.

(Middle–Lower Devonian) correspond to more distal This Tmax value is relatively high, and the maturity level
deposits, therefore classic marine type II kerogen charac- of this sample is not certain. Nevertheless, Rock-Eval
teristics have been chosen for the modeling. pyrolysis of old source rocks, such as Devonian source
rocks in Algeria, commonly show high Tmax values on
Retama Formation immature samples (C. Ducreux, 1993, personal commu-
nication). Optical data show a vitrinite reflectance value
The shaly beds of the lower Retama Formation of of 0.5%, which confirms the immaturity of this sample.
Carboniferous age, also called Toregua Formation, show On this basis, we use the computed kinetic parameters
reasonably good potential. Deposition started in the late for the modeling.
Famenian (360 Ma) and may have continued to the
Visean (327 Ma). The Toregua Formation, corresponding Copacabana Formation
to siliciclastic deposits of a marine platform, is an alterna-
tion of shaly and sandy beds which form the potential The Copacabana Formation of Permian age is a
reservoirs. This formation is present in the Madre de marine platform succession formed by an alternation of
Dios basin and in the northern sub-Andean belt (Figure shallow water (50-m) limestone, shale, and sandstone.
5C). It is about 500 m thick. Deposition began in the Stephanian (307 Ma) and
Retama source rock potential has been proven in the continued to the Early Permian (270 Ma). It is present in
Pando-X1 well (Madre de Dios basin). The total organic the Madre de Dios basin, in the northern sub-Andean
carbon (TOC) content exceeds 2 wt. % over a 100-m belt, and to a lesser extent in the central zone (Figure 5D).
interval, with a hydrogen index (HI) of more than 600 Where it has not been eroded, it is up to 800 m thick.
mg HC/g. All of the shaly beds have some potential Rock-Eval pyrolysis confirms the petroleum potential
(TOC > 1 wt. %). of this formation. Recorded TOC values vary from 1 to 9
Kinetic parameters have been determined on a sample wt. %, and the maximum HI is about 440 mg HC/g.
from the Pando-X1 well (HI = 660 mg HC/g and Tmax = Unfortunately, we have insufficient data to define the
445˚C) using OPTKIN software (Espitalié et al., 1985). true thickness of the active source rock interval.
450 Baby et al.

Figure 5—Geographic distrib-


ution of source rocks in the
Bolivian sub-Andean basins.
(A) Silurian, (B) Devonian,
(C) Toregua Formation
(Upper Devonian–Lower
Carboniferous), and (D)
Copacabana Formation
(Upper Carboniferous–Lower
Permian). Shading indicates
source rock distribution in
sub-Andean basin. SC, Santa
Cruz.

MODELING an episode of Middle Triassic extension that is known in


the southern part of Bolivia and is related to the initiation
Because all the source rocks are older than the Andean of Gondwana fragmentation (Oller and Sempere, 1990;
orogen, the key factor for exploration is the relative timing Soler and Sempere, 1993). An age of 235 Ma is used in
between maturation-migration and the formation of struc- our modeling.
tures. Any early migration of hydrocarbons would Two prominent nondepositional unconformities
exclude the young structures as drillable prospects. occur in the Cretaceous (144–68 Ma) and Paleocene
(53–27 Ma) stratigraphic sections (Figures 3, 4).
Kinematic and Geometric Hypotheses
Method
The structures have been studied on the basis of three
balanced cross sections constructed from field studies A one-dimensional model (GENEX software) has
and subsurface data (Figure 2). Original data and discus- been used to determine maturity level in real and ficti-
sion of these sections are presented elsewhere (Baby et tious wells. GENEX permits computation of the burial
al., 1989, 1992, 1993, in press). Nondepositional unconfor- history, the compaction based on porosity-depth rela-
mities and erosion surfaces recognized in the field and tionships, the temperature history, and the maturation
from seismic data have been dated as part of the YPFB- and expulsion history.
ORSTOM program (Sempere, 1990; Oller, 1992). The two Thrusting is not directly modeled by GENEX, so we
main erosional unconformities are in the Triassic and the used a fast rate of sedimentation to simulate thrust
upper Oligocene (Figures 3, 4). The latter is due to the emplacement. This approach adequately describes the
beginning of Andean deformation dated at 27 Ma burial and compaction history, but it also results in a
(Sempere et al., 1990). The age of the earlier erosion is not misleading thermal profile for a stratigraphic section
well defined, but it is believed to have occurred during spanning a few million years. The temperature of the
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 451

Figure 6—Oil and gas windows in the Lliquimuni balanced cross section. (A) Present day and (B) pre-Andean deformation.

deposited sediment is assumed to be constant during the and default marine type II characteristics for the Lower–
sedimentation process, in contrast to the overthrust Middle Devonian and Silurian. The maturation of this
sediments, which were already heated. There are two kerogen is slower than a classic type II kerogen (e.g.,
ways to overcome this problem. One can either impose a Toarcian of the Paris basin in GENEX), resulting in an
steady-state solution or artificially increase the surface estimated oil window 800 m deeper than the default
temperature to simulate a temperature closer to that of value for a gradient of about 25°C/km.
the overthrust sediments. With GENEX software, the expelled quantities are
When a steady-state solution is imposed, the relax- calculated as a function of the saturation in the source
ation time is zero. The error resulting from active rock. An expulsion threshold is defined by the user
thrusting is mainly a function of the thrust thickness and (Forbes et al., 1991). When the saturation is less than this
the rate of thrusting. The relationship of this error to the value, there is no expulsion. An arbitrary value of 15%
rate of thrusting has been calculated (Endignoux and has been used in this study.
Wolf, 1990), showing that the error is negligible for
normal rates of horizontal displacement. In this study,
the uncertainties about the timing of thrust emplacement
are so large (a few million years) that it would be inap- RESULTS
propriate to use a more sophisticated approach.
Because thermal data are scare, an average heat flow Northern Sub-Andean Zone: Lliquimuni
of 55 mW/m2 has been used for the modeling. No rifting Cross Section
phase has been included. Neither do we have enough
data to describe heat flow variations through time. A Geometry
constant value for the basal heat flow is thus used. There The northern part of the sub-Andean zone is underex-
are also uncertainties regarding the amount of Triassic plored, with only four wells having been drilled. One of
erosion as well as the burial history. These constraints, these (LQM-X1) is included on the balanced cross section
especially for Silurian and Early Devonian maturation, (Figure 6) and another is located on the Boya trend, 50
limit us to a qualitative interpretation. km south of the section. The 235-Ma erosional event
Maturation has been calculated using kinetic para- appears to have stripped up to 800 m of section. Late
meters derived from the Pando well (type Retama) for Oligocene erosion (27 Ma) was more limited; 100 m of
the Toregua Formation and the Upper Devonian section erosion is inferred for the missing section.
452 Baby et al.

Figure 7—Maturity in the Pelado structure, based on Rock-Eval pyrolysis and the subsidence history of a fictitious well,
Pelado-2 (see Figure 6 for location). (A) Subsidence history and hydrocarbon windows. (B) Calculated and measured
hydrogen index. (C) Calculated and measured Tmax values. (D) Transformation ratio of the various source rocks.

Subsidence and deposition of the deep Tertiary geometry of the thrust structures shows a later phase of
piggyback basin of the Alto Beni syncline (Figure 2A) deformation. We attribute this out-of-sequence propaga-
was controlled by Andean deformation. Seismic data tion phase to a 2.5-Ma event, which corresponds approxi-
show two unconformities within this basin fill. On the mately to the age of deposition of the conglomeratic
eastern and western flanks of the Lliquimuni anticline, Tutumo Formation. Farther east, deformation was
the sedimentary cover is up to 6000 m thick (Figure 6). younger, from 6 Ma to the present.
The unconformities within the Miocene have not been
dated, but they are known to be related to the thrusting Petroleum System
phases of the Lliquimuni and Pelado structures. For The Silurian stratigraphy in our cross section is
modeling purposes, we have used the ages of the known only west of the CFP (Figure 2A) and was not
Tertiary formations and intra-Miocene unconformities of modeled. The other three source rocks (Devonian
the southern sub-Andean zone (Marshall and Sempere, Tomachi and Tequeje formations, Toregua Formation,
1991; Gubbels et al., 1993), which we have correlated to and Copacabana Formation) are present to the west (see
the northern sub-Andean area. Ages of 11 and 6 Ma are Figure 5); eastward the Permian and Carboniferous
inferred for the Miocene unconformities (Figure 3). The sections are missing.
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 453

Figure 8—Maturity in the foreland of the Lliquimuni cross section, based on the subsidence history of a fictitious well, Eva-
Eva 1 (see Figure 6 for location). (A) Subsidence history and hydrocarbon windows. (B) Transformation ratio of the various
source rocks.

Calibration was undertaken on the Pelado structure Devonian source rock matured during this period. The
for which geochemical data (Tmax and HI) are available second phase of hydrocarbon generation was initiated by
(Figure 7). The maturation index, especially for the Andean thrusting during the late Oligocene and affected
Paleozoic source rocks, remains low along the western all the source rocks. Burial maturation is attributed to
anticlines. On the Pelado structure, Rock-Eval pyrolysis thick accumulations in the Alto Beni piggyback basin
has shown that the Copacabana Formation is immature and in the sub-Andean foreland basin during the
(Figures 7B, C), suggesting that it never was buried Neogene.
under Tertiary sediments. Farther east, seismic data Up to 4500 m of Neogene sediments were deposited
record a Tertiary section that is uniform in thickness as prior to deformation in the eastern region. The rate of
well as facies. These data suggest that the Pelado sedimentation during the Neogene was very high. The
structure developed as an early thrust that formed the thermal transient effects resulted in deep oil windows
orogenic front and carried the Alto Beni piggyback basin. (about 5.4 km depth) and late maturation.
Subsequently, the orogenic front migrated toward the
east where the Miocene cover occurs in new structures
such as the Sierra Chine, Fatima, and Eva-Eva (Figure 6).
Northern Sub-Andean Zone:
Figure 7A shows that in the Pelado structure, the two
Isiboro Cross Section
Devonian source rocks matured during the Permian. Geometry
Triassic erosion was not substantial and maturation
The only petroleum play in this section is the Isiboro
persisted to the present. The Copacabana and Toregua
frontal anticline. Other structures were deeply eroded
formations are still immature. In the piggyback basin
due to the large amount of shortening (58%) that
westward of the structure, these two source rocks
occurred in this area. Two wells have been drilled in the
matured during Andean deformation due to the thick
Isiboro anticline, and one (SSA-X1) is included in the
synorogenic deposits.
balanced cross section (Figure 9). Compared to the major
For comparison, Figure 8 shows the maturity in a ficti-
Triassic erosion, the late Oligocene unconformity was
tious well (Eva Eva-1) located on the foreland where only
relatively insignificant. The amount of Triassic erosion
the Lower and Middle Devonian source rocks are
increased from southwest to northeast, where up to 700
present. The maturity appears to have occurred in the
m of section were removed.
Neogene and is attributed to foreland infilling. The oil
window was reached during Charqui deposition (6 Ma). Petroleum System
The transformation ratio of the Tequeje Formation is now
about 70% (Figure 8B). The Copacabana, Retama, and Devonian Tomachi
The complete balanced cross section with the present and Tequeje source rocks are believed to have the same
oil windows is shown in Figure 6. There were two characteristics as their northern counterparts. The
phases of hydrocarbon generation. The first preceded Silurian with marine affinities is included as a potential
Andean deformation and began during the Carbonif- source rock interval (see Figure 5). Eastward, the
erous and affected the thick Devonian sections. Only the Permian and Carboniferous intervals are missing.
454 Baby et al.

Figure 9—Oil and gas window in the Isoboro balanced cross section. (A) Present day and (B) pre-Andean deformation.

Figure 10 shows a comparison between the maturity the Upper Cretaceous section was thinned to about 300
along the edge of the external part of the preorogenic m. Triassic erosion was substantial and removed at least
basin (well IS. 1) and along its internal part where the 2 km of Paleozoic deposits.
Devonian section is thick (well IS. 3).
Two phases of hydrocarbon generation occurred. Petroleum System
Figure 9 shows the source rocks that were matured The Copacabana and Toregua formations are missing
before and during Andean deformation. Maturation of (Figure 5). We used two source rocks in our model: the
the Silurian and Devonian sections started during the Devonian (Limoncito and Iquiri formations) and the
Devonian in the internal part where the Paleozoic is very Silurian. Rock-Eval data from the Santa Rosa X2 well
thick (up to 4000 m). Toward the east, burial resulted record a Tmax of about 440°C in the Limoncito Formation,
from foreland deposition that was contemporaneous which corresponds to the beginning of the oil window for
with Andean deformation. As in the Lliquimuni cross this type of organic matter (Retama kinetic parameters).
section, the oil window is relatively deep (up to 5 km) The subsidence and maturity measured and calculated on
due to the large amount of sedimentation. the San Juan X2 well are shown in Figure 12.
Figure 11 summarizes these maturation results on the
Central Sub-Andean Zone: Boomerang balanced cross section, showing present-day and pre-
Cross Section Andean oil windows. Even before Andean deformation,
the Silurian and Lower–Middle Devonian source rocks
Geometry were mature. Maturation started at the beginning of the
This cross section spans the principal hydrocarbon Carboniferous and expulsion was completed by the end
province of Bolivia. It is well constrained by subsurface of the Carboniferous. No structural traps are known for
seismic data and by the Santa Rosa X2 and San Juan X2 this period, thus limiting exploration opportunities to
wells (Figure 11). The Boomerang zone is formed by the stratigraphic plays. These source rocks are now in the gas
deformed border of the Paleozoic sedimentary wedge. window. During the Andean deformation, only the
Well data permit quantification of late Oligocene erosion; Upper Devonian section entered the oil window.
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 455

Figure 10—Maturity in the Isoboro cross section, comparing (A) fictitious well IS. 1 in the foreland with a thin Paleozoic
section to (B) fictitious well IS. 3 in the hinterland with a thick Paleozoic section. See Figure 9 for locations of wells.
456 Baby et al.

Figure 11—Oil and gas windows in the Boomerang balanced cross section. (A) Present day and (B) pre-Andean deformation.

FINAL GAS TO OIL RATIO development of a prominent transfer zone. To the north,
the thrusts are wider and the amount of shortening
The present proven reserves show that the sub- increases. The western part of the northern sub-Andean
Andean zone, and especially the Boomerang area, are zone is characterized by a thick Tertiary piggyback basin
mainly gas prone. All the major fields (Carasco, Katari, fill.
Palometa, Santa Rosa, Sirari, Vivora, and Yapacani) Geochemical analysis documents four source rocks in
produce gas and condensate. Nevertheless, the source the Paleozoic sedimentary wedge: the Retama Formation
rocks are of marine origin and not gas prone. We (Upper Devonian–Lower Carboniferous), the Copaca-
attribute this to the very low initial potential of the source bana Formation (Upper Carboniferous–Lower Permian),
rock. Any oil generated has not yet been expelled but the Devonian deposits (410–360 Ma), and probably the
rather remains in the source rock where burial and matu- Silurian section (430–410 Ma). Specific kinetic parameters
ration continue. Lighter compounds are formed and for the Retama Formation show a later maturation
migrate into the Andean structures. compared to other marine type II source rocks.
Our modeling emphasizes two phases of hydrocarbon
generation. The first occurred from the Devonian to
CONCLUSIONS Carboniferous due to deepening of the Paleozoic basin.
Hydrocarbons expelled before the Triassic are believed
In the northern and central sub-Andean zone of to have filled stratigraphic traps. Because of the absence
Bolivia, the propagation of the orogenic front was guided of structures and widespread erosion, most of these
by the northern boundary of the Paleozoic sedimentary hydrocarbons have probably been lost. The second phase
wedge. In the Boomerang area, this boundary is oriented of maturation is attributed to burial by Tertiary deposi-
obliquely to the regional shortening and it controlled the tion in the foreland and piggyback basins.
Petroleum System of the Northern and Central Bolivian Sub-Andean Zone 457

Figure 12—Subsidence and maturity in the San Juan X2 well (see Figure 11 for location). (A) Burial history and hydrocarbon
windows. (B) Hydrogen index versus depth.
458 Baby et al.

Acknowledgments This study resulted from a research Sempere, T., G. Hérail, J. Oller, and M. G. Bonhomme, 1990,
convention between YPFB, ORSTOM, and IFP. We thank C. Late Oligocene–early Miocene major tectonic crisis and
Ducreux (BEICIP) for the Rock-Eval analyses and kinetic related basins in Bolivia: Geology, v. 18, p. 946–949.
parameters and J. L. Pittion (Total) for the geochemical results Sheffels, B., 1988, Structural constraints on crustal shortening
in the Bolivian Andes: Ph.D. dissertation, Massachusetts
interpretation. GENEX is an IFP commercial product
Institute of Technology, Cambridge, 170 p.
marketed by BEICIP. We thank Jim Maloney, Mike Perkins, Sheffels, B., 1990, Lower bound on the amount of crustal
and an anonymous reviewer for their comments on an earlier shortening in the central Bolivian Andes: Geology, v. 18,
version of this manuscript. p. 812–815.
Soler, P., and T. Sempere, 1993, Stratigraphie, géochimie et
signification paléotectonique des roches volcaniques
basiques mésozoïques des Andes Boliviennes: Comptes
REFERENCES CITED Rendus de l’Académie des Sciences de Paris, 316, Série II,
p. 777–784.
Baby, P., G. Hérail, J. M. López, O. López, J. Oller, J. Pareja, T.
Sempere, and D. Tufiño, 1989, Structure de la zone
Subandine de Bolivie: influence de la géométrie des séries
sédimentaires antéorogéniques sur la propagation des
chevauchements: Comptes Rendus de l’Académie des
Sciences de Paris, Série II, 309, p. 1717–1722.
Baby, P., G. Hérail, R. Salinas, and T. Sempere, 1992,
Geometric and kinematic evolution of passive roof
duplexes deduced from cross section balancing: example
from the foreland thrust system of the southern Bolivian
subandean zone: Tectonics, v. 11, p. 523–536. Authors’ Mailing Addresses
Baby, P., B. Guillier, J. Oller, G. Herail, G. Montemurro, D.
P. Baby
Zubieta, and M. Specht, 1993, Structural synthesis of the
Bolivian Subandean zone, in International Symposium on ORSTOM
Andean Geodynamics: Orstom, Série “Colloques et Sémi- Laboratoire de Géodynamique des Chaînes Alpines
naires,” p. 159–162. 15, rue Maurice Gignoux
Baby, P., M. Specht, J. Oller, G. Montemurro, B. Colletta, and J. 38031 Grenoble
Letouzey, in press, The Boomerang-Chapare transfer zone France
(recent oil discovery trend in Bolivia): structural interpreta-
tion and experimental approach, in F. Roure, ed., Special I. Moretti
Publication, EAPG Congress, Moscow: Paris, Editions IFP, BP 311
Technip. 92305 Rueil Malmaison
Endignoux, L., and S. Wolf, 1990, Thermal and kinematic
France
evolution of thrust basins: a 2D numerical model, in J.
Letouzey, ed., Petroleum and tectonics in mobile belts:
Paris, Editions Technip, p. 181–192. B. Guillier
Espitalié, J., G. Deroo, and F. Marquis, 1985. La pyrolyse ORSTOM
Rock-Eval et ses applications: revue de l’Institut Français A.P. 17 11 06596
du Petrole: v. 40, p. 563–579 and 755–784. Quito
Forbes, P., P. Ungerer, A. Kuffus, F. Riis, and S. Enggen, 1991, Ecuador
Compositional modeling of petroleum generation and
expulsion: AAPG Bulletin, v. 75, p. 873–893. R. Limachi
Gubbels, T. L., B. L. Isacks, and E. Farrar, 1993, High-level E. Mendez
surfaces, plateau uplift, and foreland development,
YPFB
Bolivian central Andes: Geology, v. 21, p. 695–698.
Marshall, L. G. ,and T. Sempere, 1991, The Eocene to Pleis- C.P. 1659
tocene vertebrates of Bolivia and their stratigraphic Santa Cruz
context: a review, in Fósiles y Facies de Bolivia (1) : Revista Bolivia
Técnica de YPFB (Santa Cruz), v. 12 p. 631–652.
Oller, J., 1992, Cuadro chronoestratigrafico de Bolivia: unpub- J. Oller
lished YPFB document, Santa Cruz, Bolivia. Petrolex
Oller, J., and T Sempere, 1990, A fluvio-eolian sequence of C.P. 3969
probable Middle Triassic–Jurassic age in both Andean and Santa Cruz
Subandean Bolivia, in International Symposium on Bolivia
Andean Geodynamics: Orstom, Série “Colloques et Sémi-
naires,” p. 237–240.
Roeder, D., 1988, Andean-age structure of Eastern Cordillera M. Specht
(Province of La Paz, Bolivia): Tectonics, v. 5, p. 23–39. Total
Sempere, T., 1990, Cuadros estratigráficos de Bolivia: prop- Cedex 47
uestas nuevas: Revista Técnica de YPFB (Santa Cruz), v. 11, 92509 Paris-La Defense
p. 215-227. France
Structural Styles and Petroleum Occurrence in the
Sub-Andean Fold and Thrust Belt
of Northern Argentina

H. J. Belotti

L. L. Saccavino

G. A. Schachner
Perez Companc
Neuquén, Argentina

Abstract

T he sub-Andean fold and thrust belt of northern Argentina is characterized by a set of north-northeast
and south-southwest oriented structural trends. Major oil and gas accumulations occur in the eastern
sub-Andean ranges, including the Ramos, Aguaragüe, Campo Duran, and Madrejones fields. Thin-skinned
deformation involves a tectonic wedge that generated triangle zones bounded by a lower detachment in the
Silurian Kirusillas Formation and by an upper detachment at the base of the Upper Devonian Los Monos
Formation. In the Ramos and Aguaragüe structures, this configuration creates a distinct disharmony between
the tectonic wedge and its passive roof. The shortening is accommodated by major tectonic thickening of the
Los Monos Formation, while the overlying Carboniferous and Tertiary units are deformed into box folds. In
the Aguaragüe structure, these overlying units are cut by imbricate thrusts that offset the box fold crest
eastward with respect to the anticlinal axis of the underlying tectonic wedge. The Campo Duran structure
represents an early stage of this process without active folding of its passive roof. In contrast, the Madrejones
structure transferred shortening to the foreland, forming the Ipaguazu structure. Balanced cross sections were
used in the interpretation of structural styles and hydrocarbon traps to resolve the structures and reduce the
exploration risk.

Resumen

L a faja plegada y fallada subandina del norte de Argentina está caracterizada por un conjunto de estruc-
turas orientadas norte-noreste a sur-suroeste. Los principales yacimientos de hidrocarburos descubiertos
se ubican en su sector oriental, incluyendo los yacimientos de Ramos, Aguaragüe, Campo Duran y
Madrejones. La deformación de lámina delgada involucra una cuña tectónica que generó zonas triangulares,
con despegue inferior en la Formación Kirusillas (Silúrico) y despegue superior en la base de la Formación Los
Monos (Devónico Superior). En las estructuras de Ramos y Aguaragüe esta configuración genera una marcada
disarmonía entre la cuña tectónica y su techo pasivo. El acortamiento es resuelto por un fuerte espesamiento de
la Formación Los Monos, mientras que las unidades suprayacentes carboníferas y terciarias, son deformadas
en pliegues tipo cajón. En la estructura de Aguaragüe, estas últimas unidades son afectadas por corrimientos
imbricados que desplazan hacia el este la cresta del pliegue somero con respecto al eje anticlinal de la cuña
tectónica infrayacente. La estructura de Campo Duran representa una etapa temprana en el proceso de este
modelo, sin plegamiento activo de su techo pasivo. En contraste, la estructura de Madrejones transfiere su
acortamiento hacia el antepaís, formando la estructura de Ipaguazu. Secciones estructurales balanceadas
fueron usadas en la interpretación del estilo tectónico, y entrampamiento de hidrocarburos, como así también
resolución de estructuras y reducción de riesgos exploratorios.

Belotti, H. J., L. L. Saccavino, and G. A. Schachner, 1995, Structural styles and petroleum 545
occurrence in the sub-Andean fold and thrust belt of northern Argentina, in A. J.
Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG
Memoir 62, p. 545–555.
546 Belotti et al.

INTRODUCTION
The sub-Andean fold and thrust belt is a deformed
belt that continues from the border between Peru and
Ecuador into northern Argentina in a northwest-
southeast to NNE-SSW direction (Figure 1). In northern
Argentina, it covers an area of nearly 11,000 km2 and is
characterized by thin-skinned deformation that extends
from the Eastern Cordillera in the west to the Chaco
plain in the east (Figure 2).
The Eastern Cordillera formed in response to
compressive stress from the west that also involved the
crystalline basement (Mingramm et al., 1979) in a thick-
skinned fashion (see Vann et al., 1986). Part of the short-
ening was transmitted toward the sub-Andean fold and
thrust belt along a detachment level located at the base of
the Silurian Kirusillas Formation, and part was accom-
modated by back-thrusts at the front of the Eastern
Cordillera (Figure 3).
The Chaco plain (Russo et al., 1979) is located to the
east of the sub-Andean ranges. It represents the foreland
that was not affected by the Andean deformation.
The fold and thrust belt in Argentina is divided into
two sectors according to structural and petroleum char-
acteristics. The western sub-Andean ranges are more Figure 1—Location map of the sub-Andean ranges in
structurally complex and in an early stage of hydro- South America and their relationship to other major
tectonic elements.
carbon exploration. Only minor hydrocarbon reserves
have been discovered in Carboniferous and Tertiary
units in the Baja de Oran range (Figure 2). The eastern
sub-Andean ranges are less complex and in a slightly
more advanced stage of exploration. The main hydro- lithologic contrast determines the conspicuous sub-
carbon discoveries have occurred in Devonian, Carbonif- Andean structural style (Aramayo Flores, 1989). The
erous, and Tertiary reservoir rocks. diversity of names for the Devonian units has been
During the 1950–1970 period, Yacimientos Petrolíferos summarized by Acevedo (1986) who correlated seismic
Fiscales (YPF), the Argentinian national oil company, and well data between northern Argentina and southern
discovered significant hydrocarbon accumulations in Bolivia.
each trend. Many of these fields are large in size. The aim The Carboniferous and Triassic units are separated by
of this study is to describe the structural style and mode strong erosional unconformities and include mainly
of hydrocarbon entrapment of the principal fields in the continental and proximal shelf deposits (Stark et al.,
eastern sub-Andean ranges. Well and seismic data and 1992b; Stark, 1995). These units do not have pronounced
surface geology were used to construct balanced cross mechanical contrasts and consequently behaved
sections, some subsequently verified through seismic passively during folding.
modeling. In the Cretaceous, all of these units were affected by
regional uplift and erosion related to the formation of the
Lomas de Olmedo rift basin (Moreno, 1970; Bianucci et
GEOLOGIC FRAMEWORK al., 1981; Gómez-Omil et al., 1989; Di Persia et al., 1991;
Comínguez and Ramos, 1995). Figure 5 shows that
The stratigraphic succession in the sub-Andean ranges Paleozoic units were eroded progressively deeper
comprises Paleozoic, Mesozoic, and Tertiary marine and toward the south and east. As a result, Tertiary sedimen-
continental sedimentary rocks. The total sedimentary tary beds (Tranquitas and Chaco formations) overlie
column is about 10,000 m thick. The basal Ordovician Triassic rocks of the Cuevo Group at the Bolivian border
units were not involved in the Andean deformation and Silurian rocks of the Kirusillas Formation at the
(Figure 4). The lowest deformed Paleozoic rocks are southern end of the sub-Andean ranges (Figures 6, 7).
Silurian–Devonian units comprising alternating sand- Erosion of the Paleozoic sequence limited the southward
stones and shales. Basal proximal sandstones and development of the thin-skinned sub-Andean ranges as
conglomerates attributed to storm-dominated processes the lower detachment level within the Kirusillas
in a shallow marine basin are overlain by a pelitic shelf Formation became eroded and compressional stresses
succession. The base and the top of the sequence are inverted the basement-seated extensional faults of the
formed by angular to regional unconformities (Bottcher northwestern Argentinian rift system (Bianucci et al.,
et al., 1983; Stark et al., 1992a; Stark, 1995). This distinct 1982; Grier, 1990).
Structural Styles and Petroleum Occurrence, Sub-Andean Fold and Thrust Belt, Northern Argentina 547

Figure 2—Principal features


of the Argentinian sub-
Andean ranges. They are
divided into a western and an
eastern belt on the basis of
structural style and
petroleum characteristics.
Cross section A–A' is shown
in Figure 3.

Figure 3—Cross section


A–A' showing how the
eastward-verging Eastern
Cordillera transferred the
major part of its shortening
to the sub-Andean fold and
thrust belt. Some shortening
was accommodated as a
frontal back-thrust. See
Figure 2 for location.
(Modified after Aramayo
Flores, 1989.)

PETROLEUM GEOLOGY major hydrocarbon fields discovered in this area are


closely related to the presence of the Los Monos
Source Rock Formation. This same relationship exists in the western
sub-Andean ranges (Figure 2) where oil seeps and minor
The Los Monos Formation, composed of a succession hydrocarbon fields occur wherever this formation is
of dark gray shales and siltstones, constitutes the present.
principal generating source rock of the sub-Andean area
(Mombru and Aramayo Flores, 1986; Aramayo Flores, Reservoirs and Seals
1989). Approximately 1000 m thick, it contains two zones
of high fluid pressure that controlled hydrocarbon Main reservoirs in the sub-Andean fold and thrust
migration: the lower zone is in the lower third of the Los belt occur in the Devonian Huamampampa Formation
Monos Formation and the upper zone is close to its top. and the Carboniferous Tupambi Formation (Figure 4).
The lower zone caused migration of gas and condensate The large Ramos and Aguaragüe fields produce gas and
toward the underlying Huamampampa Formation, condensate from silica-cemented quartz arenites of the
while the upper zone caused the migration of oil or gas Huamampampa Formation, with matrix porosities of
and condensate toward the Carboniferous and Tertiary about 5% and permeabilities of 0.2 md (millidarcy)
units. (Mombru and Aramayo Flores, 1986). Major hydro-
The areal distribution of the Los Monos Formation in carbon productions are related to fractures that link
the eastern sub-Andean ranges is shown in Figure 8. The matrix porosities along the structures. Fluvial and deltaic
548 Belotti et al.

Figure 4—Lithologic column


showing stratigraphic units,
petroleum occurrences, and
detachment zones. Tertiary
section is not completely
represented.

Figure 5—Subcrop map at


the Cretaceous unconfor-
mity, showing progressive
southward erosion of the
Paleozoic succession.
Thrust faults: 1. San
Antonio, 2. Aguaragüe, 3.
Campo Duran–Madrejones,
4. Ipaguazu. Cross section
Y–Y' is shown in Figure 6,
and X–X' in Figure 7.
(Modified after Di Persia et
al., 1991.)

sandstones of the Tupambi Formation produce gas and Seal rocks include shales of the Los Monos Formation,
condensate in the Campo Duran and Madrejones fields. shales and diamictites of the Tarija Formation, and shales
Porosities average 13% and permeabilities 17 md. This in the Las Peñas and Tranquitas formations (Figure 4).
formation thins and becomes tight southward of these
fields. Secondary sandstone reservoirs of the Tarija, Las
Peñas, and Tranquitas formations produce oil in the Traps and Timing
Tranquitas, Lomitas, and Lomitas Bloque Bajo fields. The
silica-cemented quartz arenites of the Santa Rosa Traps are predominantly structural. Hydrocarbon
Formation produce gas and condensate in the Ramos accumulations are related to relative positions within the
field. thrust sheet, such as anticlinal crests, faulted anticlines,
Structural Styles and Petroleum Occurrence, Sub-Andean Fold and Thrust Belt, Northern Argentina 549

Figure 6—Schematic cross


section Y–Y' beneath basal
Miocene unconformity. See
Figure 5 for location.

Figure 7—Tracing of a seismic section east of the fold and thrust belt, showing major unconformities (solid lines) bounding
six Paleozoic–Tertiary sequences. See Figure 5 for location.

and faulted vertical fold limbs. However, combined STRUCTURAL STYLES OF THE EASTERN
stratigraphic and structural traps are possible in Tertiary
and Carboniferous sedimentary rocks related to fluvial
SUB-ANDEAN RANGES
sandstone deposits. The eastern sub-Andean ranges comprise a set of
Based on the Tertiary stratigraphic record, the devel- elongated structures that trend NNE-SSW, forming the
opment of the fold and thrust belt began in the late San Antonio, Aguaragüe, Campo Duran–Madrejones,
Miocene or early Pliocene (Ahlfeld and Branisa, 1960; and Ipaguazu ranges (Figure 8). The principal hydro-
Mingramm et al., 1979; Martínez, 1980; Jordan and carbon reserves are located in the Ramos, Aguaragüe,
Alonso, 1987). New 40Ar-39Ar dates demonstrate signifi- Campo Duran, and Madrejones fields. Smaller fields
cant upper crustal shortening within the Eastern include the San Pedrito, Macueta, Tranquitas, Lomitas,
Cordillera after 10 Ma. In the adjacent sub-Andean Lomitas Bloque Bajo, and Ipaguazu (see Figures 8, 11).
ranges, the deformation also occurred after 10 Ma To determine the structural style, we constructed a
(Gubbels et al., 1993), confirming a link between both series of balanced cross sections (Figure 8) using well
morphostructural units (Figure 3). During Andean defor- data, seismic lines, and surface geology (see Dahlstrom,
mation, hydrocarbons migrated toward the structures. 1969; Suppe, 1983; Woodward et al., 1985; Mitra and
550 Belotti et al.

Figure 8—Eastern sub-


Andean ranges and the
principal gas and condensate
fields (shaded areas). Cross
sections A–A' through D–D'
are shown in Figures 9, 11,
13, and 16, respectively.

Namson, 1989; Dahlstrom, 1990). The balancing method generates a fault-bend fold (see Suppe, 1983) with a lower
we applied consisted of conservation of bed length of the detachment in the Kirusillas Formation and an upper
competent units (e.g., Santa Rosa and Huamampampa detachment at the base of the Los Monos Formation
formations). The less competent successions (e.g., Los (Figure 9). The tectonic wedge verges eastward causing
Monos Formation) were areally balanced. Several of delamination of the sedimentary units (Figure 10). This
these cross sections were seismically modeled to check major sole thrust has a total shortening of 5.5 km, which is
their validity and correspondence with the available transferred to the overlying units by means of the upper
seismic data. detachment level. The sedimentary sequence that overlies
The Ramos and Aguaragüe ranges have high relief the tectonic wedge is folded into a box fold geometry. A
and steep surface dips. As a result, seismic profiles zone of complex deformation, occurring in a zone of high
provide data on their western flank and across the fluid pressure within the Los Monos Formation, is
synclines between ranges. In the Campo Duran, present in the core of the fold. This results in a coinci-
Madrejones, and Ipaguazu ranges, seismic data are better dence between the axis of the box fold at the surface and
because of the gentler topography and rather low dips at the crest of the tectonic wedge at depth.
the surface (see Figures 13, 16). The large Ramos field produces gas and condensate in
fractured quartz arenites of the Santa Rosa and
Huamampampa formations within the tectonic wedge.
RAMOS RANGE The oil–water contact is believed to be similar in both
formations, thus questioning the sealing capacity of the
The Ramos structure is the southernmost morpho- Icla Formation (Figure 9).
structural unit of the San Antonio range (Figure 8). The San Pedrito and Macueta fields are located to the
Others to the north are the San Pedro and Macueta struc- north of Ramos field and produce oil from the Carbonif-
tures. The trend is about 70 km long in Argentina and erous Tupambi Formation. In Bolivia, the San Alberto
extends into the San Alberto structure in Bolivia. field produces small amounts of oil from the Tarija
The Ramos structure has an asymmetric configuration Formation (Quevedo Velasco, 1977), and a recently
that is limited to the north by an oblique ramp that drilled deeper exploratory well discovered large
generated a steeply dipping closure. To the south, the amounts of gas and condensate in the Devonian
structure is gently dipping and ends near the extensional Huamampampa Formation.
faults that bound the Cretaceous Lomas de Olmedo rift To the south of Ramos, the Cretaceous unconformity
basin (Figure 5). The oldest outcropping sedimentary truncates the Los Monos Formation (Figures 5, 6),
sequence is the Tertiary Tranquitas Formation. The resulting in a change of the structural style. Here, the
structure is subdivided into two main areas on the basis upper detachment is in the Tertiary Tranquitas
of structural style. In the northern area, the structure Formation, and the shortening to the upper units is trans-
consists of a triangle zone (see Gordy et al., 1977; Jones, ferred as a fault-propagation fold to the eastern Tomasito
1982; Woodward et al., 1985) where the tectonic wedge range (Figure 8).
Structural Styles and Petroleum Occurrence, Sub-Andean Fold and Thrust Belt, Northern Argentina 551

Figure 9—Balanced structural


cross-section A–A' through
the Ramos field in the
northern part of the Ramos
range. See Figure 8 for
location.

structure (Figure 11). There is an important difference,


however: several minor imbricate thrusts that detach
near the top of the Los Monos Formation break through
the box fold geometry of the overlying Carboniferous
and Tertiary units and offset the crest of the box fold
eastward with respect to the crest of the underlying
tectonic wedge.
Wells drilled on the flanks of the structure encounter a
normal thickness of the Los Monos Formation (Figure
11). However, wells drilled through the core penetrate
up to three times the normal thickness of the Los Monos
Formation in a zone of complex deformation and high
fluid pressure.
The tectonic wedge is formed by fault-bend folding of
Figure 10—Theoretical model of tectonic wedge delami- alternating massive and thin-bedded units. These thin-
nating sedimentary units. (Modified after Allmendinger, bedded units, such as the Kirusillas, Icla, and the base of
1987.) the Los Monos Formation, required interstratal slip for
concentric folding in ramp anticlines. The massive quartz
arenites of the Huamampampa and Santa Rosa forma-
AGUARAGÜE RANGE tions adjusted to the shortening at the top of the ramp by
extensional fracturing in the fold crest and frontal limb
The Aguaragüe range is formed by a single contin- (Figure 12). This kind of trap described by Jones (1987)
uous 250-km-long structure, oriented NNE-SSW. One combines shalier units (source and seal rocks of the Los
third of it lies in Argentina and the rest in Bolivia. The Monos Formation) with massive units (fractured reser-
oldest exposed sedimentary rocks occur in the Bolivian voirs of the Huamampampa and Santa Rosa formations)
sector and belong to the Upper Devonian Los Monos at its leading edge. This is also true for the Aguaragüe
Formation (Ahlfeld and Branisa, 1960; Pareja et al., 1978). field where the best gas and condensate production is
In Argentina, the Carboniferous Tupambi Formation is located at the fold crest and the eastern flank (Grieco,
the oldest exposed unit, cropping out near the Bolivian 1991).
border (Aramayo Flores, 1989). Crustal shortening The size of the hydrocarbon accumulations in the
increases northward from 5.3 km in the Aguaragüe field Aguaragüe range is closely related to the location of the
in Argentina to 11 km near Villamontes in Bolivia. At the traps within the structure. The large Aguaragüe field
surface, the Aguaragüe range continues along strike with produces gas and condensate from fractured quartz
gentle inflections of its axis and without displacement arenites of the Huamampampa Formation within the
transfers. Like the Ramos range, its southern limit is tectonic wedge. Its distinguishing characteristic is its
marked by an oblique ramp near the extensional faults of structural continuity along strike (Mombru and
the Lomas de Olmedo Cretaceous basin (Figure 5). Aramayo Flores, 1986), reflecting the continuity of the
The structural style of the Aguaragüe structure shows major thrust associated with the tectonic wedge. The
a triangle zone geometry similar to that of the Ramos smaller Tranquitas, Lomitas, and Lomitas Bloque Bajo
552 Belotti et al.

Figure 11—Balanced structural


cross section B–B' through
Aguaragüe field. See Figure 8
for location. (Based on surface
data from Mombru and
Aramayo Flores, 1986.)

Compared to the Ramos and Aguaragüe ranges, the


Campo Duran structure represents an early stage in
triangle zone formation (Figure 13). The tectonic wedge
in this structure has a displacement of 1.5 km in the
south and 2.4 km in the north, which is less than in the
Ramos and Aguaragüe ranges. Along strike, an increase
in sole thrust shortening was accompanied by an
increase of shortening on minor thrusts formed within
the Los Monos Formation, resulting in a higher structural
elevation of the Carboniferous and Tertiary units above
Figure 12—Alternation of massive and thin-bedded units the detachment.
within the tectonic wedge creating a hydrocarbon trap Incipient back-thrusts originated east of the sole thrust
above a ramp anticline. The massive quartz arenites of the as shortening transferred out of the Campo Duran
Huamampampa and Santa Rosa formations developed a
structure (Figure 13). To the north, displacement on this
fractured reservoir at its leading edge. (After Jones, 1987.)
back-thrust increased as the minor thrust in the Campo
Duran structure died out. As a result, a syncline in the
fields produce oil from Carboniferous and Tertiary units Carboniferous and Tertiary units separates the Campo
and are deformed by minor thrusts that form shallow Duran structure from the Madrejones structure (Figure 8).
anticlines and imbricates with vertical fold limbs (Figure The large Campo Duran field produces gas and
11). Formed by many small faults, these anticlinal traps condensate from Carboniferous sandstones of Tupambi
do not extend far along strike. Formation. It consists of southeast-northwest oriented
fluvial channels and deltaic distributary mouth bars that
vary in thickness as the result of an incised valley at the
top of the Los Monos Formation.
CAMPO DURAN–MADREJONES RANGE Figure 14 is traced from a seismic profile. It was used
The Campo Duran–Madrejones range is the eastern- to construct the structural cross section of Figure 13. It
most component of the Argentinian fold and thrust belt illustrates the divergence between the hanging wall and
that involves Silurian and Devonian units. Almost the the footwall that defines the location of the sole thrust.
entire range lies within Argentina, with only the north- Although a domal shape of the Ordovician, Silurian, and
ernmost part extending into Bolivia. The northern and Devonian sequences beneath the thrust is evident on this
southern ends are marked by oblique ramps. The upper tracing, seismic modeling (Figure 15) of the Campo
Tertiary Chaco Formation is the oldest outcropping Duran structural cross section shows that these
stratigraphic unit. This range is subdivided into two sequences are not involved in the deformation and are
main areas on the basis of structural style: Campo Duran subhorizontal. The pull up in the footwall results from
to the south and Madrejones to the north (Figure 8). higher interval velocities of the overlying Paleozoic rocks
Structural Styles and Petroleum Occurrence, Sub-Andean Fold and Thrust Belt, Northern Argentina 553

Figure 13—Balanced struc-


tural cross section C–C'
through the Campo Duran
field. See Figure 8 for
location.

Figure 14—Tracing of seismic section used as the basis of Figure 15—Seismic model used to construct the Campo
the Campo Duran structural cross section (Figure 13). Duran structural cross section (Figure 13).
Velocity pull-up affects the Devonian, Silurian, and Ordovi-
cian sequences beneath the sole thrust.

in the thrust sheet compared to the lower velocities of the Monos Formation that is known as the Ipaguazu
Tertiary units in the adjacent synclines. A similar pull up structure (Figure 16). Total shortening is about 2.7 km
exists in the Aguaragüe range (Lesta and Kozlowsky, and decreases northward into Bolivia.
1992). The large Madrejones field produces gas and conden-
To the north, the Madrejones structure is a fault-bend sate from the Carboniferous Tupambi Formation. It
fold that has the same basal detachment level as the involves a broad anticlinal crest with some minor thrusts
Campo Duran structure. The difference between the two affecting the Carboniferous and Tertiary units. To the
is that shortening on the sole thrust in the Madrejones east, the Ipaguazu field contains noncommercial hydro-
structure is transferred to the foreland, forming an carbons in the Tupambi Formation on the western
eastward-verging fault-propagation fold above the Los faulted flank of the structure (Figure 16).
554 Belotti et al.

Figure 16—Balanced struc-


tural cross section D–D'
through the Madrejones and
Ipaguazu fields. See Figure 8
for location.

CONCLUSIONS 5. The large Ramos and Aguaragüe fields produce


gas and condensate from silica-cemented quartz
1. The eastward-verging Eastern Cordillera arenites of the Huamampampa and Santa Rosa
responded to compressive stress in a thick-skinned Formations. Hydrocarbon productions are related
fashion. Part of the shortening was transmitted to fractures that link matrix porosities along the
toward the thin-skinned sub-Andean fold and structures. Fluvial and deltaic sandstones of
thrust belt along a detachment located at the base Tupambi Formation produce gas and condensate
of the Silurian Kirusillas Formation, and part was in the Campo Duran and Madrejones fields.
accommodated by back-thrusts at the front of the 6. Major hydrocarbon accumulations are found in
Eastern Cordillera. This deformation occurred in sedimentary sequences that are involved in fault-
the late Miocene or early Pliocene. bend folds overlying major step thrusts. Only small
2. The southward development of the thin-skinned hydrocarbon fields are associated with minor
sub-Andean ranges was limited as the lower thrust folds that form shallow and narrow anti-
detachment in the Kirusillas Formation became clines.
eroded by the Cretaceous unconformity and as 7. The hydrocarbon fields in the eastern sub-Andean
compressional stresses inverted the basement- ranges are closely related to the presence of the Los
seated extensional faults of the northwest Argen- Monos Formation. In the western sub-Andean
tinian rift system. ranges, oil seeps and minor hydrocarbon fields
3. The sub-Andean ranges are characterized by two occur where this formation is present.
structural styles. First, triangle zones are present
where tectonic wedges generate fault-bend folds by
ramping up from a lower detachment in the Acknowledgments We would like to thank the Exploration
Kirusillas Formation to an upper detachment at the Department of Naviera Perez Companc for support and
base of the Los Monos Formation. The sedimentary permission to publish this study; Herman Welsink, Peter Jones,
sequence that overlies the tectonic wedges is folded and Gustavo Vergani for their helpful suggestions; Patricia
into a box fold. A zone of complex deformation Moretti and Diana Valenzuela for the Spanish–English trans-
associated with zones of high fluid pressure in the lation; and Silvia González for typing. Ramón Colombres,
Los Monos Formation occupies the cores of these Gustavo Pascal, and Silvia Colonna drafted the figures. Part of
folds. Second, shortening on the major thrust is this study was made during our employment with Yacimientos
transferred to the foreland where it formed an Petrolíferos Fiscales.
eastward-verging fault-propagation fold.
4. Within the tectonic wedge, the thin-bedded units,
such as the Kirusillas, Icla, and basal Los Monos
Formation, required interstratal slip for concentric REFERENCES CITED
folding in ramp anticlines. The massive quartz Acevedo, O. M., 1986, El Precarbónico en la Provincia de
arenites of the Huamampampa and Santa Rosa Salta: Boletín de Informaciones Petroleras, Buenos Aires,
formations adjusted to the shortening at the tops of Argentina: Tercera Epoca, no. 6, p. 65–72.
ramps by extensional fracturing in the fold crests Ahlfeld, F., and L. Branisa, 1960, Geología de Bolivia: La Paz,
and frontal limbs. Bolivia, Instituto Boliviano del Petróleo, 245 p.
Structural Styles and Petroleum Occurrence, Sub-Andean Fold and Thrust Belt, Northern Argentina 555

Allmendinger, R. A., 1987, Técnicas modernas de análisis Lesta, P. J., and E. Kozlowski, 1992, Las acumulaciones de gas
estructural: Asociación Geológica Argentina, Buenos Aires, y condensado en los niveles profundos del Devónico de las
Serie B, Didáctica y Complementaria, no. 16, 90 p. Sierras Subandinas: Ciencia y técnica, Buenos Aires, p.
Aramayo Flores, F., 1989, El cinturón plegado y sobrecorrido 16–26.
del norte argentino: Boletín de Informaciones Petroleras, Martínez, C., 1980, Structure et évolution de la Chîne Hercyni-
Buenos Aires, Tercera Epoca, no. 17, p. 2–16. enne et de la Chaîne Andien dans le nord de la Cordellére
Bottcher, G., M. Frigerio, N. Samosiuk, and M. Vistalli, 1983, des Andes de Bolivie: Travaux et documents de I’Orstom,
Sedimentitas devónicas en el subsuelo de la Sierra Baja de v. 119, 352 p.
Oran. Estratigrafía e intento de caracterización paleoambi- Mingramm, A., A. Russo, A. Pozzo, and L. Cazau, 1979,
ental: Revista Técnica Yacimientos Petroliferos Fiscales Sierras Subandinas, in Segundo Simposio de Geología
Bolivianos, La Paz, no. 9, p. 1–4. Regional Argentina: Córdoba, Argentina, Academia
Bianucci, H. A., O. M. Acevedo and J. J. Cerdán, 1981, Nacional de Ciencias, v. 1, p. 95–138.
Evolución tectosedimentaria del Grupo Salta en la Mitra, S., and J. S. Namson, 1989, Equal-area balancing:
subcuenca de Lomas de Olmedo (provincias de Salta y American Journal of Science, v. 289, p. 563–599.
Formosa): Noveno Congreso Geológico Argentino, Actas Mombru, C., and F. Aramayo Flores, 1986, Geología del
3, San Luis, p. 159–172. yacimiento Aguaragüe: Boletín de Informaciones Petrol-
Bianucci, H., J. F. Homovc, and O. M. Acevedo, 1982, eras, Buenos Aires, Tercera Epoca, no. 6, p. 53–64.
Inversión tectónica y plegamientos resultantes en la Moreno, J. A., 1970, Estratigrafía y paleogeografía del
comarca Puesto Guardián-Dos Puntitas, departamento Cretácico Superior en la cuenca del noroeste argentino, con
Orán, Provincia de Salta: Primer Congreso Nacional de especial mención a los subgrupos Balbuena y Santa
Hidrocarburos, Petróleo y Gas, Exploración, Buenos Aires, Bárbara: Revista Asociación Geológica Argentina, Buenos
p. 23–30. Aires, v. 25, p. 9–44.
Comínguez, A. H., and V. A. Ramos, 1995, Geometry and Pareja, J., C. Vargas, R. Suárez, R. Ballón, R. Carrasco, and C.
seismic expression of the Cretaceous Salta rift system of Villarroel, 1978, Mapa Geológico de Bolivia y memoria
northwestern Argentina, in A. J. Tankard, R. Suarez, and explicativa: La Paz, Bolivia, Yacimientos Petrolíferos
H. J. Welsink, Petroleum basins of South America: AAPG Fiscales Bolivianos y Servicio Geológico de Bolivia, scale
Memoir 62, this volume. 1:1,000,000, 2 sheets.
Dahlstrom, C. D. A., 1969, Balanced cross sections: Canadian Quevedo Velasco, J., 1977, Geología de los yacimientos de
Journal of Earth Sciences, v. 6, p. 743–757. hidrocarburos en Bolivia: Revista Técnica de Yacimientos
Dahlstrom, C. D. A., 1990, Geometric constraints derived from Petrolíferos Fiscales Bolivianos, La Paz, Bolivia, v. 1-2,
the law of conservation of volume and applied to evolu- p. 47–121.
tionary models for detachment folding: AAPG Bulletin, v. Russo, A., R. Ferello, and G. Chebli, 1979, Llanura Chaco
74, p. 336–334. Pampeana: Geología Regional Argentina, Academia
Di Persia, D. E., R. J. Carle, and H. Belotti, 1991, Geología Nacional de Ciencias, Córdoba, v. 1, p. 139–183.
petrolera en la subcuenca de Lomas de Olmedo: Boletín de Starck D., F. Gallardo, and A. Schultz, 1992a, La discordancia
Informaciones Petroleras, Buenos Aires, Tercera Epoca, no. precarbónica en la porción argentina de la cuenca de
25, p. 14–29. Tarija: Boletín de Informaciones Petroleras, Buenos Aires,
Gómez-Omil, R. J., A. Boll, and R. Hernandez, 1989, Cuenca Tercera Epoca, no. 29, p. 2–11.
cretácica-terciaria del noroeste argentino (Grupo Salta): Starck, D., E. Gallardo, and A. Schulz, 1992b, La cuenca de
Cuencas sedimentarias argentinas, Universidad de Tarija: estratigrafía de la porción argentina: Boletín de
Tucumán, Argentina, v. 6, p. 43–64. Informaciones Petroleras, Buenos Aires, Tercera Epoca, no.
Gordy, P. L., F. R. Frey and D. K. Norris, 1977, Geological 30, p. 2–14.
guide for the Canadian Society of Petroleum Geologists Starck, D., 1995, Silurian–Jurassic stratigraphy and basin
and 1977 Waterton–Glacier Park Field Conference: evolution of northwestern Argentina, in A. J. Tankard, R.
Canadian Society of Petroleum Geologists, Calgary, 93 p. Suarez, and H. J. Welsink, Petroleum basins of South
Grieco, L. F., 1991, Distribución de fluidos en la estructura de America: AAPG Memoir 62, this volume.
Aguaragüe: Boletín de Informaciones Petroleras, Buenos Suppe, J., 1983, Geometry and kinematic of fault-bend folding:
Aires, Tercera Epoca, no. 26, p. 61–66. American Journal of Science, v. 283, p. 684–721.
Grier, M. E., 1990, The influence of the Cretaceous Salta rift Vann, I. R., R. H. Graham, and A. B. Hayward, 1986, The
basin on the development of Andean structural geome- structure of mountain fronts: Journal of Structural
tries, NW Argentine Andes: Ph.D. dissertation, Cornell Geology, v. 8, p. 215–227.
University, Ithaca, New York, 178 p. Woodward, N. B., S. E. Boyer, and J. Suppe, 1985, Balanced
Gubbels, T. L., B. L. Isacks, and E. Farrar, 1993, High-level geological cross sections: American Geophysical Union,
surfaces, plateau uplift, and foreland development, Short Course, Washington, D.C., v. 6, 170 p.
Bolivian central Andes: Geology, v. 21, p. 695–698.
Jones, P. B., 1982, Oil and gas beneath east-dipping under-
thrust faults in the Alberta foothills, Canada, in R. B. Authors’ Mailing Addresses
Powers ed., Geologic studies of the Cordilleran thrust belt, H. J. Belotti
v. 1: Denver, Rocky Mountain Association of Geologists,
L. L. Saccavino
p. 61–74.
Jones, P. B., 1987, Quantitative geometry of thrust and fold
G. A. Schachner
belt structures: AAPG Methods in Exploration Series, no. 6, Perez Companc
26 p. J.J. Lastra Sud 6000
Jordan, T. E., and R. N. Alonso, 1987, Cenozoic stratigraphy C.C. 181
and basin tectonics of the Andes Mountains, 20°–28° south 8300 Neuquén
latitude: AAPG Bulletin, v. 71, p. 49–64 Argentina
Northern South America

MERIDA ANDES of western Venezuela, with frailejón shrubs in foreground.


The Merida range is a striking example of pronounced vertical uplift that
occurred along basement faults during the Andean orogeny, forming a giant
“pop-up” structure.

Edgar Ortiz, 1994, watercolor, 30 × 23 cm


Northern Part of Oriente Basin, Ecuador:
Reflection Seismic Expression of Structures

H. R. Balkwill F. I. Paredes

G. Rodrigue J. P. Almeida

Petro-Canada Resources, Tripetrol Petroleum Ecuador


Calgary, Alberta, Canada Quito, Ecuador

Abstract

I ndustry reflection seismic profiles from the northern part of the Oriente basin display families of
basement-rooted structures ranging in age from early Mesozoic (and possibly Permian) to Quaternary. We
interpret the Mesozoic-Cenozoic structures to be kinematically and chronologically compatible with tectonic
events displayed in the contiguous Andean Cordillera. Late Triassic (and Permian?) extensional structures may
have been linked to an intra-Cordilleran rift regime. Widely developed Late Jurassic-Early Cretaceous conver-
gent structures are coeval with transpression along the western margin of the South American plate. Late
Cretaceous and Cenozoic convergent structures are responses to major episodes of plate marginal terrane
accretion and plate convergence. Late Cretaceous and Cenozoic phases of structuring are displayed on seismic
profiles as a network of steeply dipping northward-trending faults that have risen from Precambrian crystalline
basement into the Phanerozoic cover rocks. Northward-elongated sharply hinged folds generated in cover
rocks by slip on basement faults are traps for Oriente basin oil fields.

Resumen

L os perfiles de reflexión sismica disponibles dentro de la industria de la parte norte de la cuenca del
Oriente, muestran varios typos de estructuras asociadas con el basamento que datan desde el Mesozoico
superior (y posiblemente del Pérmico) haste el Cuaternario. Las estructuras del Mesozoico-Cenozoico se inter-
pretan cinematica y cronológicamente compatible con eventos tectónicos que se manifiestan adyacentes a la
Cordillera Andina. Tales eventos son los siguientes: Las estructuras de extensión del Triasico Superior pueden
ester asociadas a un régimen de "rift" intra-cordillera. Las estructuras convergentes, ampliamente desarrolladas
al Jurasico Superior-Cretacio Inferior, son contemporaneas con la compresión a lo largo del margen oeste de la
place suramericana. Las estructuras convergentes del Cretacio Superior y del Cenozoico son resultados debido
a mayores episodios de acreción de terrenos y de convergencia de la place. Las differentes etapas de desarrollo
de las estructuras del Cretacio Superior y del Cenozoico se manifiestan como un conjunto de fallas de fuerte
buzamiento de dirección norte, que surgen del basamento cristalino Precambrico hasta la cobertura Phanero-
zoica. Los pliegues agudos de type tumbados de direccion norte, generados en la cobertura por deslizamiento
de fallas al nivel del basamento, son trampas pare los campos de la cuenca del Oriente.

INTRODUCTION southward into Peru, where it is called the Marañon basin.


These domains are part of the system of sub-Andean
Ecuador can be divided into three distinctive subpar- foreland basins that reach from Venezuela to southern
allel tectonic-morphologic provinces, which from east to Chile (Gansser, 1973).
west are the Oriente, Cordillera (Andes), and Costa Cumulative oil production from the Ecuadorian
province (Baldock, 1982) (Figure 1). The jungle-covered Oriente has been more than 1.5 billion bbl; current
Oriente basin is the focus of on-going petroleum explo- production is about 300,000 bbl per day. Oriente basin oil
ration. The Oriente basin extends northward into is housed in Cretaceous sandstones, trapped in structures
Colombia where it is called the Putamayo basin, and of Cretaceous and Tertiary ages (White et al., 1995).

559
Balkwill, H. R., G. Rodrigue, F. I. Paredes, and J. P. Almeida, 1995, Northern part of Oriente basin,
Ecuador: reflection seismic expression of structures, in A. J. Tankard, R. Suárez S., and H. J. Welsink,
Petroleum basins of South America: AAPG Memoir 62, p. 559-571.
560 Balkwill et al.

Figure 1--Regional tectonic


map of the Oriente basin,
Ecuador. Note locations of
seismic sections in Fgures
3-12. (Adapted from
Baldock, 1982; Rosania and
Morales, 1986; and from
mapping of regional seismic
grid by Petroecuador and
Petro-Canada.)

Geochemical analyses indicate that the oil migrated into quality ranges from poor to excellent; the quality and
these structures from Cretaceous source rocks in areas coverage are adequate for regional mapping purposes.
now comprising the eastern Cordillera and southern- Tschopp's (1953) paper on early oil exploration in the
most parts of the Oriente basin (Dashwood and Abbotts, Oriente region provides a standard valuable reference
1990). Rivadeneira (1988) suggested that Lower Jurassic for stratigraphic nomenclature and early concepts of
source shales in southern Ecuador and northern Peru Oriente geology. Canfield et al. (1982) have provided an
may have provided some of the oils. informative paper on the geology of the large Sacha oil
Petroleum exploration progressed generally eastward field in the western part of the basin (Figure 1).
and southward across the Oriente basin. Large, topo- Dashwood and Abbotts (1990) have summarized Oriente
graphically expressed anticlines of the sub-Andean basin petroleum geology, with an emphasis on the
foothills were drilled in early operations (Tschopp, 1953). geochemistry of oils and source rocks. Abundant unpub-
Present exploration is directed mainly to subtle struc- lished geologic and geophysical reports have resulted
tures concealed beneath the Oriente jungle (Canfield et from work by the staff of Petroecuador (formerly Corpo-
al., 1982; Dashwood and Abbotts, 1990). More than 550 racion Estatal Petrolera Ecuatoriana, or CEPE).
wells have been drilled, most of which bottomed in This paper provides representative reflection seismic
Mesozoic rocks. A few wells penetrated through the sub- profiles across the northern part of the Oriente basin to
Mesozoic unconformity to Paleozoic strata and to illustrate the styles and ages of oil-bearing and prospec-
Precambrian crystalline basement. In excess of 30,000 km tive structures and to provide evidence for interpreta-
of CDP reflection seismic data have been acquired since tions of the Mesozoic-Cenozoic tectonic environment of
1961. The average seismic grid is about 4 x 4 km, but it is the basin. Much of the material and concepts resulted
denser in some areas. Data have been recorded and from a collaborative regional study by CEPE and Petro-
processed usually to 5-6 sec (two-way time). Seismic Canada explorationists in 1987-88 (Balkwill et al., 1988).
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 561

REGIONAL SETTING and Phanerozoic strata lie under the moderately incised
sub-Andean domain. These rocks are largely correlative
By reason of its great elevation (locally higher than 5000 with the less disturbed rocks of the eastern part of the
m), the Cordillera forms the dominant physiographic basin. Two anticlinoria (Napo and Cutucu uplifts)
province of Ecuador (Gansser, 1973) (Figure 1). The moun- separated by an intervening structural depression (Puyo
tainous Cordillera consists of two subparallel belts depression) dominate the structural geometry of the sub-
separated by a narrow graben. The Western Cordillera has Andean domain (Figure 1). Westward-dipping steep
a basement of Upper Cretaceous oceanic crust that was thrusts, which bring ruggedly incised Cordilleran crys-
obducted onto the margin of South America in latest talline rocks against the unmetamorphosed sub-Andean
Cretaceous-early Tertiary time and subsequently restruc- rocks, form an abrupt western boundary for the sub-
tured by Tertiary plate dynamics (Daly, 1989). The Eastern Andean foothills.
Cordillera (Cordillera Real) consists of an infrastructure of
polydeformed, variably metamorphosed Mesozoic and
older rocks, with a superstructure of Cenozoic andesitic STRATIGRAPHY
flows, volcaniclastics, and associated sedimentary rocks.
Jaillard et al. (1990) recognized an initial Mesozoic The Oriente basin cover rocks range in age from
tectonic event (possibly as old as Permian) in the region of Silurian to Quaternary; regional lithostratigraphic, bio-
the Cordillera Real. It involved Triassic continental margin stratigraphic, and reflection seismic evidence allows delin-
rifting, linked possibly northward through Colombia to eation of several megasequences (Figure 2). The succession
the vast Tethyan rift domain of the proto-Caribbean and thickens westward and southward to more than 5 km at
mid-Atlantic. The Triassic rift domain was superposed in the border with Peru (Figure 1).
the Early Jurassic by a marine trough (Santiago trough) in Only a few wells have reached crystalline basement
which carbonates, shales, and subaerial volcanic rocks under the Oriente foreland basin cover. From this meager
accumulated. From detailed mapping of outcrops in the well control and from regional mapping of shallow reflec-
Cordillera Real, Aspden and Litherland (1992) proposed a tion seismic character, the basement under the entire
tectonic scenario for subsequent Jurassic-Cretaceous Oriente is believed to consist of Proterozoic metamorphic
development of the Cordillera Real. They interpreted the and plutonic rocks of the Amazon craton (Baldock, 1982;
western margin of the pre-Mesozoic craton to lie near the Canfield et al., 1982). As a generalization, radiometric ages
eastern edge of the Cordillera, where large-scale calc- of Amazon craton rocks decrease systematically
alkaline volcanic and plutonic activity commenced in the westward, outlining northwest-trending geochronologic
late Early Jurassic and continued to middle Late Jurassic. provinces that may represent terranes affixed successively
Structural fabrics mapped in the Cordillera by Aspden to an Archean core during the early-middle Proterozoic.
and Litherland (1992) indicate dextral transform and The youngest province recognized thus far (Sunsas mobile
eastward overthrusting along steeply dipping shear zones. belt) may represent a phase of ensialic deformation,
The shear zones bound distinctive tectonic-lithologic metaphorphism, and plutonism during the interval 1100-
assemblages ("divisions"), which they interpreted as 900 Ma (Teixeira et al., 1989; deMatos and Brown, 1992).
allochthonous terranes, accreted to the margin and Regional gravity data indicate that the sialic crust under
deformed from the latest Jurassic to middle Early Creta- the Oriente basin is about 30-35 km thick (Feininger and
ceous (about 125 Ma). In addition, they correlated uplift of Seguin, 1983).
the Cordillera Real and major resetting of mineral ages in Paleozoic rocks in the sub-Andean foothills are
the middle Late Cretaceous-early Tertiary (85-55 Ma) with assigned to the Pumbuiza and Macuma formations
convergence generated by accretion of the allochthonous (Tschopp, 1953). From these outcrops and some deep
oceanic rocks now comprising the basement of the wells, the Pumbuiza terrigenous clastics have yielded
Western Cordillera and Costa (Figure 1). marine fossils ranging in age from Late Silurian to Early
Neogene plate reorganization in the eastern Pacific Carboniferous. The Macuma carbonates and clastics
caused accelerated convergence at the northwestern contain fossils ranging in age from Late Carboniferous to
margin of South America, resulting in great uplift of the Permian (Canfield et al., 1982).
Cordillera, a surge of andesitic volcanism, and develop- The Pumbuiza and Macuma formations cannot be
ment of present-day Andean morphology (Herron, 1972; easily separated on most seismic profiles (Figures 3,4,5,6).
Daly, 1989). There is sufficient regional control to outline a western
The Oriente basin province consists of two physio- region of Pumbuiza-Macuma rocks under the proximal
graphic-structural domains. A topographically low, part of the Oriente foreland and an eastern domain along
jungle-covered eastern region has a floor of Precambrian the Ecuador-Peru border. The two domains of Paleozoic
crystalline basement, overlain by a westward-thickening strata are separated by an irregularly shaped northward-
wedge of slightly folded and faulted Phanerozoic cover elongated region of Precambrian basement (Figure 1). The
rocks. A system of large, westward-dipping reverse faults Paleozoic strata typically lie in isolated structures bounded
(and locally an eastward-facing steep monocline) by basement-rooted faults that display large-scale reversal
separates the eastern Oriente subprovince from a of dip slip (Figures 4, 5, 6). The combined thickness of
contiguous sub-Andean segment. Moderately to strongly Pumbuiza-Macuma beds in some of these structures is
folded and faulted eastward-displaced cratonic basement greater than several hundred meters (Rosania and
562 Balkwill et al.

Figure 2--Stratigraphy of the northern part of the Oriente basin. (Adapted from Canfield et al., 1982; Rosania and Morales,
1986; Petro-Canada proprietary reports.)
Morales, 1986). recovered from some well samples (M. Rivadeneira,
The basement fault-bounded structures that define Petroecuador, 1988, personal communication). Regional
inliers of Pumbuiza-Macuma strata commonly contain a seismic evidence from northern Peru (Marañon basin--
superimposed succession, hundreds of meters thick, of not available for display--shows that the fault-bounded
conglomeratic, nonmarine, terrigenous clastics (Figures clastic rocks are older than the Jurassic Chapiza
3, 4, 5, 6). These successions are indistinctly layered on Formation. Balkwill et al. (1988) have suggested that the
most seismic profiles. They lie in structures interpretable graben-associated beds might be broadly correlative
as post-Macuma extensional half-grabens subsequently with the Mitu Formation of the Peruvian Andes. We
modified by reversal of dip slip. The rocks were drilled now propose that the rocks represent isolated rift
at a location near seismic line CP 3530 (Figure 6). They elements that were linked westward to the cratonic
have long been considered as correlatives of the Jurassic marginal Triassic rift regime recognized recently in the
Chapiza Formation of the sub-Andean Foothills because Cordillera Real by Jaillard et al. (1990) and Aspden and
of their texture and red color and because they were Litherland (1992).
evidently younger than Macuma strata and older than Lower Jurassic marine carbonates and shales
Cretaceous (Tschopp, 1953). No firmly diagnostic fossils (Santiago Formation) crop out in the Cutucu uplift
have been reported from the enigmatic strata from this (Baldock, 1982). These rocks and coeval volcanic breccias
location or other wells where the fault-bounded beds and tuffs were deposited in an initially deep narrow
have been drilled, although possible Triassic spores were trough, superimposed on the Triassic rift domain of the
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 563

Figure 3--Part of the reflection seismic proflie CP 48 from the eastern Oriente basin (see Figure 1 for location). Vertical scale
for this and all following seismic displays is in seconds (two~way time). Rock units: PC, Precambrian basement; Pz,
Pumbuiza-Macuma formations; Tr, Triassic beds; Kh, Hollin Formation; Kn, marker in Napo Formation; Tt, top of Tiyuyacu
Formation. Note the paraconformable contact of Paleozoic strata with basement, the local truncation of Macuma beds
beneath indistinctly layered Triassic beds, and the structural disruption of Triassic and older rocks and their truncation
beneath subhorizontal basal Cretaceous strata.

Figure 4--Part of seismic section


CP 3506 (unmigrated) from the
northeastern Oriente basin.
Rock units as in Figure 3. Note
the paraconformable contact of
the Paleozoic strata (Pz) above
the basement broken by
basement-rooted extension
faults (at left). Also note the
inverted half-graben (at left) and
basement-rooted reverse fault
(at right) produced by middle
Tertiary convergence.

Cordillera Real Qaillard et al., 1990). Platform carbonates demonstrating Jurassic syndepositional uplift along the
of the upper part of the Santiago Formation are present in western margin of the Oriente basin. The clastic rocks and
the southern part of the Oriente basin. Coeval nonmarine the overlying volcaniclastic Misahualli Member (Figure 2)
strata may extend northward along the axial part of the are approximately coeval with some large Jurassic calc-
basin. alkaline plutons in the easternmost part of the Cordillera
Coarse-grained, Jurassic red beds assigned to the Real. Aspden and Litherland (1992) propose that the
Chapiza Formation in the sub-Andean foothills (Baldock, intrusive phase was followed by latest Jurassic-Early
1982) may be partially coeval with Santiago strata Cretaceous accretion and eastward thrusting of exotic
(Aspden and Litherland, 1992). The Chapiza clastic terranes against the western margin of the Cordillera
wedge tapers and becomes finer grained eastward, Real.
564 Balkwill et al.

Figure 5--Part of reflection seismic section CP 3588 from the eastern Oriente basin. Rock units as in Figure 3. Note the trun-
cation of Triassic beds beneath Cretaceous strata and the middle Tertiary inversion of Triassic extension fault. The late
Tertiary (far left) and middle Tertiary (left center) narrow contractional "pop-up" anticlines are focused at the base of
Triassic strata, possibly from intrastratal slip associated with a basement reverse fault. Moderately westward-dipping intra-
basement reflectors were activated as slip surfaces in late Cenozoic convergence. The basement-rooted late Cenozoic
reverse faults with opposed vergence define a large crustal "pop-up" and associated forced folds in the cover rocks.

Figure 6--Part of reflection seismic section CP 3530 from the northeastern Oriente basin. Rock units as in Figure 3. Note the
variably dipping intrabasement reflectors, the paraconformable contact of Paleozoic strata with basement, and the indis-
tinct reflection character of Triassic rocks.Large-scale structural disruption and truncation of Triassic beds occur beneath
basal Cretaceous Hollin Formation. Note the inversion of the Triassic half-graben by late Cenozoic contraction.
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 565

Figure 7--Part of a reflection


seismic section from the central
Oriente basin (location confiden-
tial). Rock units as in Figure 3.
Note the low-relief asymmetric
anticline in Cretaceous strata in
the central part of the section.
Also note the thinning of intra-
Cretaceous reflector "a" toward
anticlinal crest and the rejuvena-
tion of anticline in the middle
Tertiary marked by a subtle fold at
the level of Tiyuyacu Formation.

Hollin quartz arenites, possibly as old as Aptian nantly fine-grained and are a few hundreds of meters
(Aleman and Marksteiner, 1993), form the basal beds of thick. Locally prominent disconformities between the
the Cretaceous megasequence. Hollin strata lie uncon- Tiyuyacu Formation and overlying Orteguaza strata
formably above an erosionally planed surface developed mark important intrabasin structuring in Eocene or
on previously structured older rock units ranging from Oligocene time (Figure 2). The structural relief developed
Precambrian basement to Chapiza red beds (Figures 2 during this interval delineated elongate anticlinal traps in
through 12). Unpublished biostratigraphic determina- Cretaceous sandstones that host the largest oil fields in
tions (Petro-Canada proprietary information) indicate the Oriente basin (Canfield et al., 1982).
that the Hollin Formation, which is locally up to 130 m The upper terrigenous clastic sequence (Oligocene
thick, is diachronously younger eastward. The sand- Orteguaza Formation and overlying strata) represents
stones are coastal and shallow marine facies of Creta- depositional responses to accelerated convergence of
ceous clastic wedges that prograded westward into the cratonic and oceanic crust at the western margin of the
Oriente basin from its eastern margin (Figure 2). The South American plate. This was accompanied by
distal marine facies (Napo Formation) is as thick as a few vigorous uplift of the Andes and eastward-shedding of
hundred meters; it may range in age from Albian to coarse-grained clastic wedges into the contiguous Oriente
Campanian (Dashwood and Abbotts, 1990). foreland basin. Orteguaza beds lie above truncated
The eastern part of the Napo succession contains Tiyuyacu beds in the periphery of the sub-Andean belt.
southwest-tapering and southwest-fining stacked Eastward from there, the basal Orteguaza strata (which
wedges of quartz arenites and subarkoses, informally contain some thin marine beds) occupy paleodepressions
named from upper to lower, the M1, M2, U, and T sand- in the broad folds that were developed in underlying
stones. The sandstones are separated by shales and lime- Tiyuyacu beds during middle Tertiary tectonism. Large
stones (the latter also having some informal member amounts of latest Tertiary-Quaternary clastic detritus
designations). Local intraformational parasequence were shed basinward through the Puyo structural
boundaries are evident in the Napo Formation lying depression (Figure 1). The entire Cenozoic depositional
above elongate, northward-trending paleostructural cycle of eastward-prograding clastics and the older cover
highs (Figure 7). Subtle intra-Napo relief on these rocks under it are undergoing present-day structural
elements may represent local erosion and winnowing of modification by neotectonic movement on basement
strata and/or onlap by basal beds of the overlying reverse faults (Figures 5, 6, 9,10,11,12).
parasequence. Some leaching of Napo limestones at the
local unconformities indicates that the paleotopographic STRUCTURAL GEOLOGY
highs were sometimes exposed. The upper part of the
Napo Formation was truncated progressively westward Division of the Oriente basin into the sub-Andean and
from the axial part of the basin prior to deposition of eastern Oriente domains was originally based on topo-
overlying beds of the Maastrichtian - lower Tertiary Tena graphic expression of structures (Tschopp, 1953). Most of
Formation (Figure 2). the sub-Andean part of the Oriente consists of large, high,
Upper Maastrichtian-Cenozoic strata consist of coarse- northward-trending strike ridges, with outcrops of folded
to fine-grained, mainly nonmarine terrigenous clastic Paleozoic and younger strata cut by steeply dipping
detritus shed eastward in response to uplift and erosion reverse faults. The eastern Oriente is a very low, morpho-
of the Andean orogenic belt. The lower part of this logically subdued terrain where underlying structures
succession, the Tena and Tiyuyacu formations, are domi- are manifested only locally by gentle surface dips and
566 Balkwill et al.

Figure 8--Eastern part of seismic section CP 149 from the north-central Oriente basin. Rock units as in Figure 3. Note the
opposed vergence of basement-rooted reverse faults and associated "kink-band" forced folds. Fault at left is early
Tertiary age, with differential thickening of Tiyuyacu Formation (Tt), and fault at right is middle Tertiary age, associated
with thickening of post-Tiyuyacu strata.

Figure 9--Part of seismic section CP-


3506 from the northeastern Oriente
basin. Rock units as in Figure 3. Note
the eastward-dipping basement-rooted
reverse faults and forced folds in cover
rocks, having different ages of
movement: fault A, early Tertiary only;
faults B and C, Late Cretaceous or early
Tertiary, rejuvenated slightly in middle
Tertiary; fault D, prominent late
Cenozoic development, with associated
"kink-band" forced fold.
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 567

Figure 10--Part of seismic section CP 3514 from the northeastern Oriente basin. Rock units as in Rgure 3. Note the
westward-dipping late Cenozoic basement-generated "pop-up" anticline; fault at the eastern margin of structure may reach
the surface.

subtle alignments of drainage elements. influenced structures generated within those strata. In
Abundant industry reflection seismic profiles show that other places, the basement fabrics are truncated abruptly at
the most significant structures in the foreland comprise a the upper surface of the basement and seem not to
network of steeply dipping faults arranged in north- influence the cover rocks. It is inferred from this that the
trending en echelon sets (Figure 1). The faults rise from basement discontinuities originated during Proterozoic
basement into various stratigraphic levels in the cover rock events and that some of these have been reactivated--as
succession and have disrupted the cover rocks as forced innate zones of brittle failure at various times in the
(drape) folds. The strike directions of basement faults in Phanerozoic, depending on the state of stress and the
the eastern Oriente subdomain are approximately parallel nature and orientation of the basement inhomogeneity.
with structural trends in the contiguous sub-Andean Napo The genetic origins of the basement fabrics may represent
and Cutucu uplifts. Faults north of Curaray River strike large thrust faults or shear zones imparted during Protero-
north-northeastward, parallel to the general trend of the zoic accretionary events.
Napo uplift, and the foreland faults south of the Curaray Lower Paleozoic (Pumbuiza) rocks in Cutucu uplift
River strike north-northwestward, parallel to the Cutucu were folded and faulted prior to deposition of overlying
uplift. A broad inference is that the regional structural Carboniferous (Macuma) beds (Tschopp, 1953). The
trends owe their parallelism to an interconnected through- Pumbuiza and Macuma megasequences are difficult to
going deep basement fabric. Because many of the separate (except locally) in the subsurface of the northern
basement faults have undergone kinematic reversal in Oriente basin. We could not discern evidence from the
their histories, they are discussed here in a chronologic seismic data of significant middle Paleozoic structures.
framework.
Permian-Triassic
Proterozoic and Paleozoic
A network of basement-involved half-grabens and
The oldest structures evident on northern Oriente other relict structural depressions extend from the
seismic profiles are moderately dipping prominent reflec- northern Oriente basin southward to Rio Curaray, thence
tions in basement, shown in some profiles near the eastern southeastward into the Marañon basin of Peru (Figures 3,
border of Ecuador (Figures 4, 5, 6). The bundles of strong 4, 5, 6). The half-grabens commonly contain two strati-
reflectors interrupt a generally diffuse basement seismic graphic sequences. The lower succession consists of
character, commonly lacking distinctive planar reflection moderately well-stratified strong reflectors that are subpar-
characteristics. Some of the basement fabrics have struc- allel internally and with the underlying basement contact.
tural continuity into Phanerozoic cover rocks and have We interpret these rocks as carbonates and clastics of the
568 Balkwill et al.

Figure 11--Part of seismic


section CP 3550 from the
eastern Oriente basin. Rock
units as in Figure 3. Note the
prominent westward-dipping
basement-rooted reverse fault
and narrow, asymmetric "pop-
up" anticline developed by an
antithetic reverse fault rooted
near the base of Cretaceous
cover rocks.

Macuma Formation and older Pumbuiza Formation. The steep dips from crystalline basement and cut overlying
upper sequence, with its indistinctly layered, diffuse to Cretaceous-Cenozoic strata or have developed sharply
vaguely apparent internal reflection characteristics, are hinged, asymmetric forced folds with relief of a few tens
conglomeratic terrigenous clastics penetrated by the well of meters to hundreds of meters (Figures 4-12). The
drilled near seismic line CP 3530 (Figure 6). basement-generated faults and folds have critical signifi-
We interpret the extensional elements and their coarse cance for hydrocarbon exploration. All of the major fields
clastic infill to be relict cratonic inliers, genetically linked thus far discovered in the eastern Oriente basin, such as
to the Triassic rift regime that Jaillard et al. (1990) and the Sacha field (Figure 1), are low-relief structural traps,
Aspden and Litherland (1992) mapped in the Cordillera owing their existence to closure on these elements
Real and which extend along the eastern Andes of (Canfield et al., 1982).
Colombia and Peru. The oldest seismically recognizable Late Cretaceous
The early Mesozoic rift assemblage and older rocks structures are low relief, northward-elongated paleoto-
were significantly folded, uplifted, and erosionally pographic highs against which beds in the upper part of
planed prior to preservation of basal Cretaceous deposits the Napo Formation are seen to thin (Figure 7). Napo
(Figures 3, 4, 5, 6). This impressive intracratonic conver- limestones are leached at the crests of some of these
gent tectonic event reached eastward as far as the structures. We interpret these elements as small-scale
western part of Solimoes basin, western Brazil, where Late Cretaceous convergent uplifts of basement. They
deep seismic profiles show evidence of major Late appear to be embryonic foreland basin responses to the
Jurassic - Early Cretaceous reverse faulting, uplift, and Late Cretaceous transpressional convergence that
planation before deposition of the oldest preserved Aspden and Litherland (1992) interpreted for the
Cretaceous strata (deMatos and Brown, 1992; see also Cordillera Real.
Petrobras, 1983). The evidence of intra-Cordilleran trans- Differential thinning and thickening of the Tena-
pressive accretionary convergence during the Late Tiyuyacu succession, across basement-rooted reverse
Jurassic-Early Cretaceous, elucidated by Aspden and faults and forced folds, demonstrate that basin conver-
Litherland (1992), provides a possible dynamic cause for gence accelerated in the early Tertiary from mild begin-
this important intraplate event. nings in the Late Cretaceous (Figures 8, 9). The early
Tertiary foreland convergent phase was dynamically
Cretaceous-Cenozoic compatible with the regionally convergent tectonic envi-
ronment that prevailed in the Cordillera (Daly, 1989;
Cretaceous-Cenozoic representatives of the network Aspden and Litherland, 1992). A late Cenozoic surge of
of Oriente basin faults extend upward with moderate to basin compression, coincident with convergence and
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 569

Figure 12--Part of seismic section CP 149 from the sub-Andean foothills, northwestem Oriente basin. Rock units as in
Figure 3. Note the westward-dipping late Cenozoic reverse fault and splays, as well as the eastward-dipping antithetic
reverse faults.

great uplift of the Andes, is made evident by the amounts basement faults and associated folds comprise two
of structural relief generated on structures far out in the immense lobes, which meet at a structural reentrant
foreland basin (Figures 5, 6,11) as well as in the sub- crossed by the Curaray River (Figure 1) and extend from
Andean foothills (Figure 12). there westward to the Puyo depression (Figure 1). This
Some of the Cenozoic faults are up-section projections vaguely defined zone is about 30 km wide. Within it there
of faults displaying Triassic (and Permian?) normal slip. are abundant, short, variably oriented Cenozoic faults and
Slip on the Cenozoic components of these vestigial struc- associated folds. The zone separates the dominantly north-
tures has been in a reverse sense (Figures 4, 5, 6). Many northeastward structural grain of the northern Oriente
other basement faults with Cenozoic slip or forced-fold basin and Putamayo basin from the dominantly north-
flexure do not coincide with early Mesozoic (or Paleozoic) northwestward grain of the southern part of the basin and
structures (Figures 8, 9,10,11). Whether or not relict Marañon basin.
basement fabrics determined their locations cannot be We lack deep crustal reflection seismic data that might
demonstrated. The amounts and times of reverse slip and delineate the fashion in which the network of Oriente
associated forced-fold flexure vary across members of the basin basement faults are adjusted in the lithosphere.
Cenozoic network, even for faults in close proximity Focal depths for sub-Andean foreland earthquakes in
(Figure 9). For the large fault that forms the leading edge Colombia, Peru, and Ecuador are in the range of 8-38 km,
of the faulted domain north of Curaray River (Figure 1), indicating that all of the crust and possibly the uppermost
the region of greatest Cenozoic reverse slip coincides mantle are involved in deformation (Suarez et al., 1983).
approximately with the position of greatest Triassic (and At those depths, the basement faults may meet a zone of
Permian?) normal slip (Figure 6), indicating that some of intracrustal detachment or shear, possibly inherited from
the inherited basement anisotropies are particularly Proterozoic terrane fabrics, similiar in style and kinematics
susceptible to large-scale stress adjustment. East-dipping to the Andean foreland of western Argentina (Jordan and
reverse faults, antithetic to the large westward-dipping Allmendinger, 1986; Cahill et al., 1992).
reverse faults that form the leading edge of the disrupted
foreland, have created anticlinal "pop-up" culminations
that range in width from a few kilometers to many kilo-
meters (Figures 5,10). Some other east-dipping antithetic CONCLUSIONS
faults forming pop-ups are detached in ductile cover rocks
above basement rather than being rooted in basement Reflection seismic data from the northern Oriente basin
(Figure 11). display families of structures having different ages and
The regional geometry of Cretaceous-Cenozoic struc- styles that involve Precambrian crystalline basement and
tures in the Oriente foreland indicates that the network of Phanerozoic cover rocks. We interpret these cratonic struc-
570 Balkwill et al.

tural elements as products of episodic plate events along reservoirs (mainly Hollin and Napo sandstones) on
the Ecuadorian margin of the South American craton. northward-elongated, low-relief folds developed above
Episodic terrane accretion established the Amazon the basement-rooted reverse faults (e.g., Canfield et al.,
craton in the Proterozoic (Litherland et al., 1985; deMatos 1982). We have shown that some of the structures origi-
and Brown, 1992). The Oriente basin Paleozoic strata nated in the Late Cretaceous, some in the early Tertiary
were deposited on a widespread shallow marine shelf and have since been dormant, and others have rejuve-
(Baldock, 1982); the regional tectonic setting of the basin nated or moved initially in the late Cenozoic and were
during deposition of these rocks has not been elucidated. episodically rejuvenated in the Cenozoic. Critical rela-
Half-grabens containing continental red clastics may tionships may exist in terms of the ages of fold closure
be remnants of a Triassic (and Permian?) rift domain, and the times of regional oil migration and charging of
linked westward to the Gondwana rift system demon- traps.
strated in the eastern Andes of Ecuador, Colombia, and
Peru (Jaillard et al., 1990; Aspden and Litherland, 1992).
The Triassic rift elements and the Paleozoic and Precam-
brian rocks lying under and between them were reverse
faulted and uplifted during a phase of widespread Acknowledgments The writers thank Petroecuador and
intraplate convergence within the western South Petro-Canada for permission to publish this material, much of
American craton. This was possibly associated with plate which resulted from a collaborative Petroecuador/Petro-Canada
marginal transpression in the eastern Cordillera Real in regional project in 1988. We thank reviewers Norman
the Late Jurassic-Early Cretaceous. The Jurassic and Haimila, Robert Meneley, and Howard White for their
older rocks and structures were erosionally planed prior constructive comments.
to deposition of basal Cretaceous (Aptian?) Hollin sand-
stones. REFERENCES CITED
A second phase of transpressional convergence in the
Cordillera Real in Late Cretaceous-early Tertiary time Aleman, A. M., and R. M. Marksteiner, 1993, Mesozoic and
was concurrent with obduction of oceanic crust now Cenozoic tectonic evolution of the Marañon basin in south-
lying under the Western Cordillera and Costa. This eastern Colombia, eastern Ecuador and northeastern Peru
convergence prompted development in the Oriente (abs.): AAPG Bulletin, v. 77, p. 301.
basin of a network of low-relief basement reverse faults, Aspden, J. A., and M. Litherland, 1992, The Geology and
mainly having northward strikes. Propagation of the Mesozoic collisional history of the Cordillera Real,
Ecuador: Tectonophysics, v. 205, p. 187-204.
faults generated elongate folds in lower Tertiary and Baldock, J. W., 1982, Geology of Ecuador (Explanatory
older cover rocks. Late Cenozoic accelerated conver- Bulletin of the National Geological Map of the Republic of
gence of the South American and Pacific (Nazca) plates is Ecuador): Ministerio de Resursos Naturales y Energeticos,
continuing to accentuate the relief on some of the Direccion General de Geologia y Minas, Quito, Ecuador,
basement-rooted reverse faults and their associated 59p.
folds. Some small-scale intrastratal detachment may Balkwill, H. R., F. I. Paredes, and J. P. Almeida, 1988, Relation-
have taken place locally as adjustment to inversion of ships of intra-crustal delamination, stratigraphy, and oil
basement faults (Figure 5). However, we have seen no prospectivity, Oriente basin, Ecuador: Segundo Congreso
evidence in the seismic profiles from the northern Latinoamericano de Hidrocarburos, Rio de Janeiro, p. 1
Oriente basin of the large-scale, thin-skinned thrust /15-15/15.
Cahill, T., B. L. Isacks, D. Whitman, J. L. Chatelain, A. Perez,
structures seen in the sub-Andean foothills of the Llanos and J. M. Chiu, 1992, Seismicity and tectonics in Jujuy
basin of Colombia (Cooper et al., 1995). Province, northwestern Argentina: Tectonics, v. 11,p. 944-
We interpret the evidence from the regional seismic 959.
grid to indicate that the basement under the Oriente Canfield, R. W., G. Bonilla, and R. K. Robbins, 1982, Sacha
basin has a complex, anisotropic fabric consisting of large oilfield of Ecuadorian Oriente: AAPG Bulletin, v. 66,
and small basement blocks that lie above a deep, region- p.1076-1090.
ally distributed intrabasement detachment zone. This Cooper, M. A., F. T. Addison, R. Alvarez, A. B. Hayward, S.
basement fabric was probably inherited from Proterozoic Howe, A. J. Pulham, and A. Taborda, 1995, Basin develop
tectonic events. The fabric is capable of large-scale adjust- ment and tectonic history of the Llanos basin, Colombia, in
ment to regional intraplate stress regimes, whether in A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum
basins of South America: AAPG Memoir 62, this volume.
divergence or convergence. Daly, M. C., 1989, Correlation between Nazca/Farallon plate
The relatively subtle structures of the Oriente basin, kinematics and forearc basin evolution in Ecuador:
rooted in a zone of basement delamination, can be Tectonics, v. 8, p. 769-790.
viewed as embryonic examples of the types of large- Dashwood, M. F., and I. L. Abbotts, 1990, Aspects of the
scale, high-relief basement-involved foreland deforma- petroleum geology of the Oriente basin, Ecuador, in J.
tion spectacularly displayed in some other sub-Andean Brooks, ed., Classic petroleum provinces: Geological
regions, such as the Sierras Pampeanas of Argentina Society of London Special Publication, no. 50, p. 89-117.
(Jordan and Allmendinger, 1986), and in the western deMatos, R. M. D., and L. D. Brown, 1992, Deep seismic
United States (Lowell, 1977). profile of the Amazonian craton (northern Brazil):
Large oil fields in the Oriente basin are in Cretaceous Tectonics, v.11, p. 621-633.
Northern Part of Oriente Basin, Ecuador: R Reflection Seismic Expression of Structures 571

Feininger, T., and M. K. Seguin, 1983, Simple Bouguer gravity Suarez, G., P. Molnar, and B. C. Burchfiel, 1983, Seismicity, fault
anomaly field and the inferred crustal structure of conti- plane solutions, depth of faulting, and active tectonics of the
nental Ecuador: Geology, v.11, p. 40-4. Andes of Peru, Ecuador, and southern Colombia: Journal of
Gansser, A., 1973, Facts and theories on the Andes: Journal of Geophysical Research, v. 88, p. 10,403-10,428.
the Geological Society of London, v. 129, p. 83-131. Teixeira, W., C. G. Tassinari, U. G. Corani, and K. Kawashita,
Herron, E. M., 1972, Sea-floor spreading and the Cenozoic 1989, A review of the geochronology of the Amazonian
history of the east-central Pacific: GSA Bulletin, v. 83, p.1671- craton: tectonic implications: Precambrian Research, v. 42, p.
1692. 213-227.
Jaillard, E., P. Soler, G. Cartier, and T. Mourier, 1990, Geody- Tschopp, H. J., 1953, Oil explorations in the Oriente of Ecuador:
namic evolution of the northern and central Andes during AAPG Bulletin, v. 37, p. 2303-2347.
early to middle Mesozoic times: a Tethyan model: Journal of White, H. J., R. A. Skopec, F. A. Ramirez, J. A. Rodas, and G.
the Geological Society of London, v.147, p.1009-1022. Bonilla, 1995, Reservoir characterization of the Hollin and
Jordan, T. E., and R. W. Allmendinger, 1986, The Sierras Napo formations, western Oriente basin, Ecuador, in A. J.
Pampeanas of Argentina: a modern analogue of Rocky Tankard, R. Suarez, and H. J. Welsink, Petroleum basins of
Mountain foreland deformation: American Journal of South America: AAPG Memoir 62, this volume.
Science, v. 280, p. 737-764.
Litherland, M., B. A. Klinck, E. A. O'Connor, and P. E. J.
Pitfield, 1985, Andean-trending mobile belts in the Brazilian
shield: Nature, v. 314, p. 345-348. Authors' Mailing Addresses
Lowell, J. D., 1977, Underthrusting origin for thrust-fold belts
with applications to the Idaho-Wyoming belt: Wyoming H. R. Balkwill
Geological Association Guidebook, 29th Annual Field G. Rodrigue
Conference, p. 449-455. Petro-Canada Resources
Petrobras, 1983, Acre and upper Amazon basin, Brazil, in A. W. P.O. Box 2844
Bally, ed., Seismic expression of structural styles: AAPG Calgary, Alberta T2P 3E3
Studies in Geology Series, no.15, v. 2, p. 2.2.4-9 - 2.2.4-15. Canada
Rivadeneira, M., 1988, Characteristicas geological fundamen-
tales y problemas exploratoria que plantean las principales
cuencas sedimentarias ecuatorianas: Segundo Congreso
F. I. Paredes
Latinoamericano de Hidrocarburos, Rio de Janeiro, p. 1/18- J. P. Almeida
18/18. Tripetrol Petroleum Ecuador
Rosania, G., and M. Morales, 1986, Compilacion paleo- av. Gonzalez Suarez 432
geografica del Oriente Ecuatonano: Segundo Congreso Quito
Colombiano de Petroleo, Memorias Tomo 1, p. I/5-I/21. Ecuador
Reservoir Characterization of the Hollin and Napo
Formations, Western Oriente Basin, Ecuador

Howard J. White Jose A. Rodas


Oryx Ecuador Energy Company
Robert A. Skopec Quito, Ecuador
Felix A. Ramirez
Guido Bonilla
Oryx Energy Company Petroecuador
Dallas, Texas, U.S.A. Quito, Ecuador

Abstract

T he Oriente basin of Ecuador has produced a substantial amount of oil over the past 20 years. Nearly
3 billion bbl of oil have been recovered from the principal reservoirs in the Cretaceous Napo and Hollin
formations. Subtle north-south structures, commonly associated with Andean-related faulting, have trapped
much of the recoverable hydrocarbons in the thicker sandstones deposited within the Hollin and Napo reser-
voirs. East to west thinning of these reservoir units also contributes to the formation of stratigraphic traps. Both
the Hollin and Napo formations comprise successions of eastward-sourced fluvial and deltaic sedimentary
deposits that prograded westward into shoreline and marine shelf parasequences. The Albian Hollin reservoir
interval consists of a dominant alluvial plain sandstone sequence (Main Hollin sandstone) that occupies much
of the Oriente basin. In the western Oriente, the uppermost Hollin section grades vertically into open marine
strata with isolated tidal- and storm-influenced sandstone bodies. The overlying Napo stratigraphy also
consists of sand-rich fluvial and deltaic deposits in the eastern Oriente and abruptly changes to marine shales
and limestones and lowstand valley-fill sandstones in the western part of the basin. Extensive structural and
stratigraphic trap potential remains within the Napo and Hollin strata in the Oriente basin. High-resolution
geophysical techniques and detailed geologic reservoir characterization facilitate successful exploitation of
these remaining reserves.

Resumen

E n los últimos veinte años la Cuenca Oriente del Ecuador ha producido una cantidad sustancial de
hidrocarburos. Alrededor de tres mil millones de barriles de petroleo han sido recuperados de los reser-
vorios principales de las formaciones cretácicas Hollin y Napo. Estructuras sutiles orientadas norte-sur,
comunmente asociadas con fallamiento de edad Andina, han entrampado la mayoría de los hidrocarburos
recuperables dentro de los espesos depósitos arenosos de los reservorios de Napo y Hollin. La formación de
trampas estratigraficas ha estado favorecida por los adelagazamientos este-oeste de dichas unidades reservo-
rios. Las formaciones Napo y Hollin comprenden una sucesión de sedimentos deltaicos y fluviales alimen-
tados desde el este, los cuales progradaron hacia el oeste integrando parasecuencias de zonas de playa y
marino-plataformicas. El reservorio Albense Hollin consiste de una secuencia predominantemente arenosa de
planicie aluvial (Arenisca Hollin Principal) la cual se encuentra ocupando la mayoría de la Cuenca Oriente. En
el occidente del Oriente, la sección superior de Hollin grada verticalmente a sedimentos marino-abiertos con
cuerpos arenosos influenciados por mareas y tormentas. La sobreyacente estratigrafia de Napo tambien
consiste, en el este del Oriente, de depósitos deltaicos y fluviales ricos en arena, los cuales cambian abrupta-
mente a calizas y lutitas marinas, y areniscas “lowstand” de relleno de valle en la parte oeste de la cuenca.
Existe enorme potencial en trampas estructurales y estratigraficas dentro de los estratos Napo y Hollin de la
Cuenca Oriente. Las técnicas geofísicas de alta resolución y la caracterización geologica de los reservorios
facilitaran una explotación exitosa de las reservas remanentes.

White, H. J., R. A. Skopec, F. A. Ramirez, J. A. Rodas, and G. Bonilla, 1995, Reservoir char- 573
acteristics of the Hollin and Napo formations, western Oriente basin, Ecuador, in A. J.
Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG
Memoir 62, p. 573–596.
574 White et al.

Figure 1—Overview of the


Oriente basin, Ecuador,
showing structural features,
distribution of producing fields,
location of cross sections
shown in Figures 7, 15, and 16,
and specific wells referred to in
the text.

INTRODUCTION greater than 80,000 km 2. It is contiguous with the


The Oriente basin of Ecuador produces a substantial Puntamayo basin of Colombia and the Marañon basin of
amount of oil and provides attractive exploration oppor- Peru. Stratigraphically and structurally, the Oriente
tunities. The Hollin and Napo sandstone reservoirs have preserves a complex Phanerozoic geologic history
proven to be consistent producers since initial produc- beginning with earliest Paleozoic deposition and culmi-
tion was first established in August 1972. The Napo nating with Tertiary deposits shed from the Andean fold
and thrust belt. Figure 2 shows the general stratigraphy
sandstone has a cumulative production (December 1992)
of the Oriente basin. The succession is subdivided into
of 1.17 billion bbl, and the Hollin Formation has a cumu-
several unconformity-bounded sequences: Paleozoic–
lative production of 1.70 billion bbl. Production estimates
Jurassic stratigraphy, the Cretaceous Hollin and Napo
for the next 20 years are about 2 billion bbl, which will be
formations, and Upper Cretaceous–Quaternary sedimen-
derived from the currently producing fields with addi-
tary sequence.
tional reserves from fields in the process of development.
The sedimentary fill rests on Precambrian igneous
This paper presents an integrated geologic study of
and metamorphic basement that has been intersected by
the Hollin and Napo sandstone reservoirs in the greater
several wells in the eastern Oriente adjacent to the
Oriente basin, with emphasis on the western Oriente.
Guyana shield. The Silurian–Jurassic interval consists of
Figure 1 illustrates the regional setting of the Oriente
several thousand meters of carbonates, shales, and
foreland basin in front of the Andean fold and thrust
subordinate sandstones and conglomerates that have
belt, as well as the distribution of producing fields.
been structurally deformed during several episodes of
uplift and extension. Examples of the pre-Hollin
STRUCTURAL AND STRATIGRAPHIC structure are documented by Balkwill et al. (1995). The
uppermost strata of this interval belong to the Jurassic
SETTING Chapiza Formation and the associated Misahualli
The Oriente basin of eastern Ecuador is part of the volcanics. In parts of the western Oriente basin, the
upper Amazon River drainage basin and covers an area Hollin–Napo interval unconformably overlies the
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 575

eastern margin is masked by basin margin arches related


to the Guyana shield (Figure 1). Structural arches shown
on the regional map of Dashwood and Abbotts (1990)
indicate two east-west trending arches extending
westward from the Guyana shield. Interval isopach
maps confirm the existence of these two intrabasin
highs—the Aguarico arch to the north and the Cononaco
platform or arch to the south. The arches are believed to
have provided sediments as well as localized the Hollin
and Napo fluvial systems. Reservoirs within the Hollin
and Napo formations are structurally less deformed than
the underlying strata. Nevertheless, Hollin and Napo
structures are large enough to form prolific hydrocarbon
traps. A complex of major reverse faults mark the
western limit of the present-day basin adjacent to the
Napo uplift. Most of the major oil fields occur east of this
complex in structural traps that parallel the north-south
structural grain. The source rocks for these reservoirs are
believed to be the organic-rich Napo shales which have
been extensively tested for maturity, as reported by
Dashwood and Abbotts (1990).
The overlying Upper Cretaceous–Paleogene sedimen-
tary rocks in the Oriente basin were the first to be influ-
enced by incipient Andean movement. The Tena and
Tiyuyacu formations (Figure 2) are the earliest strata of
the post-Napo basin fill and consist of interbedded
shales, sandstones, and minor conglomerates. The basal
sandstone of the Tena Formation in the western Oriente
was probably derived by erosion and local reworking of
uppermost Napo. Late Tena and Tiyuyacu deposition
consisted of episodes of continental redbeds and limited
incursions of marine deposition. Deposition continued
with the Orteguaza and post-Orteguaza formations and
consisted mainly of clay-rich continental strata eroded
from the Andean volcanics. These continental sedimen-
tary rocks mark the infill of the Andean foreland basin
and comprise over 1600 m of section in the western
Oriente. The Andean volcanic arc controlled much of the
Tertiary sedimentation along the western margin of the
Oriente basin.

HOLLIN STRATIGRAPHY AND


DEPOSITIONAL SYSTEMS
Characterization of the Hollin and Napo reservoirs
includes data from seismic, well logs, core descriptions,
and petrophysical analyses. Over 1100 m of core were
examined. Regional mapping of the Ecuador Oriente
basin focused on depositional environments, paleoshore-
line trends, facies distribution, and reservoir continuity.
Figure 2—Stratigraphic column for the Oriente basin. The Hollin Formation occurs throughout the Oriente
basin. It thickens from a zero edge along the eastern
margin to nearly 200 m thick, forming a sand-rich
Jurassic Chapiza and Misahualli volcanics, but elsewhere blanket composed of several depositional sequences.
the Cretaceous rocks overlie Paleozoic strata and Figure 3 is an isopach map of Hollin strata from its
Precambrian basement. pinchout in the eastern Oriente to the depocenter in the
The Hollin–Napo interval consists of up to 500 m of southwestern part of the basin.
continental and marine sandstones, shales, and carbon- In the western Oriente basin, the Hollin can be subdi-
ates. The basin deepens toward the southwest, while its vided into the Main Hollin sandstone and the thinner
576 White et al.

Figure 3—Isopach map of the entire Hollin sequence.


Contour interval is 25 m.
Figure 4—Hollin stratigraphic column, western Oriente
Upper Hollin sandstone. General Hollin stratigraphy is basin, showing the five depositional systems comprising
described elsewhere (Wasson and Sinclair, 1927; the complete Hollin sequence.
Tschopp, 1953; Campbell, 1970; Canfield et al., 1982;
Dashwood and Abbotts, 1990; Canfield, 1991). The
Hollin is Albian in age, although the basal strata of the by isolated Hollin outcrops near Puyo. Valley fill
Main Hollin may date to the late Aptian. Faunal and thickness varies up to 40 m, reflecting the original topo-
flora taxa, although sparse, suggest that the Hollin is time graphic relief. The resulting depositional surface was a
transgressive and trace the overall sea level rise during very low relief, gently tilted surface over which braid-
late Hollin and early Napo deposition. The Hollin in the plain deposition occurred.
western Oriente basin consists of five successive deposi-
tional sequences: three sequences in the Main Hollin Braidplain Deposition
Sandstone and two in the Upper Hollin Formation The dominant depositional package in the Main
(Figure 4). Hollin consists of stacked cross-bedded sandstone and
subordinate intervals of interbedded mudstone and
Main Hollin Sandstone sandstone (Figures 4, 5) of Albian age. The upper part of
this interval forms the main oil reservoirs in the western
Valley Fill Deposition Oriente basin.
The initial Main Hollin sediments occupied substan- There are three lithofacies types in the braidplain
tial relief that was eroded into the underlying Jurassic sequence. The first and predominant one consists of
strata. Several wells penetrate this valley fill succession, stacked channel sandstones that range in thickness from
including Oso #1 and Entre Rios X1. The sedimentary 3 m to more than 10 m. The sandstones are quartzose in
rocks are interpreted to be paleovalley fluvial deposits of composition and fine to very coarse grained. Granule
channel sandstone and flood basin shales; no cores of this conglomeratic lag occasionally overlies channel scour
interval have been taken. Figure 5 shows these character- surfaces. Channel units generally have a uniform grain
istics in the Oso #1 well. This interpretation is also size distribution without any obvious fining-upward
supported by Pungarayacu cores in the Napo uplift and trend. Internally, the sandstone units are structured by
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 577

These sedimentary facies are attributed to deposition


in a braided alluvial plain environment. The overall
facies architecture resembles the Platte River and Bijou
Creek models of Miall (1977) and Cant (1982). In this
environment, the planar and trough cross bed sets are
interpreted as straight and sinuous-crested mid-channel
bars. River discharge may have been seasonally variable,
but the sand bedload was sufficiently high to account for
the stacked braid bars that dominate the Main Hollin. In
addition to sediment supply, local topography, precipita-
tion, and vegetation also influenced fluvial deposition.
Facies two, interbedded mudstone and sandstone, are
attributed to channel, overbank levee, and crevasse splay
deposition. The third facies, dominantly thick
mudstones, is inferred to have originated as channel
abandonment intervals and laterally equivalent flood
basin deposits.

Coastal Plain Deposition


The contact between the Main and Upper Hollin
formations has historically been picked on electric logs at
the base of the thicker shales overlying the stacked sand-
stones of the Main Hollin. Detailed core examination of
this contact in the western Oriente basin demonstrates
that, instead of the stacked sandstones at the top of the
Main Hollin, this sequence is frequently a fining-upward
succession of planar to trough cross-bedded sandstones
and thin mudstones. These sandstones are slightly finer
than the braided sandstones, thin upward, and are rhyth-
mically interbedded with numerous, thin, laminated
mudstones (Figure 6c). Rooted horizons are occasionally
present in the sandstones. The distinctive appearance of
these lithofacies (at least in cored intervals) indicates
Figure 5—Hollin lithofacies and depositional systems coastal plain deposition. The package ranges up to 15 m
within the Oso #1 well, Block 7, western Oriente basin. The in thickness in the western Oriente.
depositional systems have been interpreted in more than
100 Oriente wells.
The coastal plain depositional sequence is interpreted
to represent the overall abandonment of the Main Hollin
fluvial system. As such, a lower energy, higher sinuosity
planar tabular and trough cross beds in sets 30–100 cm or fluvial and estuary depositional system is envisaged for
thicker (Figure 6a). Cross bed slip faces commonly these capping sediments. The interval’s fining upward
display a grain size segregation typical of avalanche character, the occurrence of planar and trough cross
processes. Sandstone units are separated by erosion bedding and ripple lamination, the rhythmic mudstone
surfaces with carbonaceous shale laminae and mudstone interbeds, and the occasional rooting are interpreted to
beds up to 30 cm thick. Macerated plant debris often have originated in a meandering stream system that
occurs as concentrations along cross bed laminae. Dia- likely entered a coastline estuary setting.
genetic kaolinite is dispersed throughout the sandstones. The features observed in the coastal plain deposition
The resulting gamma ray signature is that of a shaly of the Main Hollin are similar to the meandering fluvial
sandstone rather than the high-porosity sandstones that to estuary profiles recognized by Smith (1987). The
typify the Main Hollin. interbedded sandstones and mudstones in the top part of
The second facies type consists of fining-upward the estuary profile exemplify features observed in tidally
channel units of finer grained sandstones and influenced sediments, although the tidal reworking is
interbedded mudstones generally less than 30 cm thick. minor (microtidal) in the coastal plain deposits.
These sandstones are more poorly sorted than the first
sandstone facies and are dominated by trough cross
bedding.
Upper Hollin Formation
Facies three comprises mudstones and mudstone
interbedded with thin sandstones. The sandstone Shore Zone Deposition
interbeds are fine to very fine grained and often ripple The upper Hollin transgressively overlies the coastal
laminated. The mudstones are massive to faintly plain veneer of the Main Hollin sandstone. It has been
laminated and often dolomitic. This facies is up to 13 m generally interpreted as a marine deposit (Dashwood
thick (e.g., Oso #1 well, Figure 5). and Abbotts, 1990; Canfield, 1991). This transgressive
578 White et al.

(a) (b)

(c) (d)

Figure 6—Core photographs from the Main and Upper Hollin formations. Photographed slabs are from conventional 10-cm
whole-diameter cores. (a) Typical planar cross bedding within the braided fluvial facies. (b) Cross-bedded sandstones and
thin mudstones of the coastal plain facies. (c) Interbedded sandstones and shales of the tidally reworked Upper Hollin shore
zone facies. (d) Open marine glauconitic sandstones beneath the capping limestones of the Upper Hollin.
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 579

blanket occurs throughout the Oriente basin, except in always interbedded with marine shelf shales (lenticular-
the extreme northeast. It consists of two distinctive litho- bedded burrowed mudstone). The quartzitic beds origi-
facies associations. The lower shore zone deposition nated as nearshore, tidally reworked marine sands. The
comprises a sandstone and shale complex that varies up glauconite-rich sands accumulated more seaward of the
to 15 m thick in the western Oriente. The upper open shoreline as storm-generated sand waves. These shoals
marine sequence caps the overall abandonment of the incorporated whatever quartz sand reached the middle
Hollin depositional system. shelf position, as well as the glauconite-replaced fecal
The shore zone lithofacies consist of fine- to medium- material derived on the shelf, and any reworked inverte-
grained planar to trough cross-bedded sandstones, very brate shell debris. It is likely that fossil abundance
fine to fine-grained ripple laminated sandstones, and increased away from the clastic shoreline. The shelf area
burrowed lenticular-bedded mudstones. Above the basal beyond the glauconitic shoals provided sites for thin but
mudstones, the shore zone locally displays a vertical widespread carbonate deposition.
profile of stacked cross-bedded and ripple-laminated The fossiliferous, micritic limestone and marls
sandstone with minor shale interbeds. A few kilometers capping the Upper Hollin record the final phase of
away, the profile may be dominantly lenticular Hollin deposition as the sea transgressed eastward over
mudstone and isolated, thin sandstones of limited the Cretaceous Oriente margin. Because of the typical
reservoir quality. The majority of the ripple sandstones thickness of the limestones (less than 2 m), seismic
contain abundant clay drapes within the lamination. The amplitude contrast at the top of the Hollin is generally
coarser sandstones occasionally exhibit strongly oblique minimal. The acoustic contrast does increase locally
cross-bedding orientations. Lenticular mudstones are where limestone thickness increases.
moderately to weakly burrowed (Chondrites, Planolites,
and minor Teichichnus) with rare ripple-laminated Pungarayacu Area
sandstone lenses.
A variety of shoreline to shallow marine depositional The Pungarayacu concession is located on the
environments combined to create the shore zone litho- basinward margin of the Napo uplift. It is noted for a
facies. A continued transgression of the tidally influ- large, shallow, heavy oil reservoir that has been
enced coastal plain resulted in the formation of sand- evaluated by Petroecuador (Almeida et al., 1983). The
dominated bay head deltas, estuaries, and subtidal Pungarayacu #2 well is representative of the entire
shoals. Muddy tidal flat and shallow marine mud depo- Hollin section in this area west of the Oriente basin. Eight
sition locally dominate the shore zone lithofacies. braided channel sandstone packages are present in the
Main Hollin. Significant mudstone intervals separate
Open Marine Deposition several of the upper channel sequences. These
The open marine succession completes the trangres- mudstones indicate periodic abandonment and aggrada-
sive Upper Hollin depositional sequence. It may range tion of the alluvial plain. The lower channels, above the
up to 15 m in thickness. The lithofacies consists of glau- Misahualli volcanics, are sand rich and devoid of
conitic and quartzose sandstone, limestone, marl, and interbedded mudstone. Figure 7 shows a north-south
shale. Ripple-laminated, very fine to fine-grained oriented cross section through three of the Pungarayacu
quartzose sandstones are commonly thin bedded and wells. Rapid facies changes occur within the distal or
moderately burrowed and occur at the base of the open westward fluvial components of the Main Hollin. The
marine succession. In the upper part of the open marine abundance of mudstone suggests that the braidplain
sequence, glauconite-rich sandstones (Figure 6d) are depositional system responsible for the Main Hollin
capped by a thin veneer of micritic and fossiliferous lime- became a mixed sand and mud system as it prograded
stones and marls. The sandstones are typically very fine westward. The Upper Hollin is represented by relatively
to fine-grained quartz with fine- to medium-grained thin mud-rich beds, subordinate quartzose and glau-
glauconite which vary up to 1 m in individual bed conitic sandstone, and capping limestone.
thickness and form sharp-based tabular units. Mudstone Hollin stratigraphy is well exposed along the
rip-up clasts occasionally overlie scour surfaces. Sedi- Hollin–Loreto road in the Pungarayacu area on the Napo
mentary structures include trough cross bedding, ripple uplift (see de Souza Cruz, 1989). A composite section of
lamination, and flaser bedding. Bioturbation often oblit- the Main and Upper Hollin strata exposed in the roadcut
erates all primary sedimentary structures. This unit is shown in Figure 8. The Main Hollin consists princi-
contains an open marine biota, including ammonites and pally of the braidplain facies, which unconformably
both thick- and thin-shelled bivalves. In the glauconitic overlies the Misahualli volcanics. The braidplain succes-
sandstones, the glauconite content is locally in excess of sion is locally saturated with oil. The outcrop shows
50% of the framework grains. Capping limestones (fossil- well-developed levee and floodplain deposits (Figure 8).
iferous wackestones) and marl beds generally measure The lower braided channel sandstones in this section are
less than 2 m thick and are well lithified. Vuggy porosity comparable with the Bijou Creek model. The capping
due to shell dissolution occurs sporatically. beds of the Main Hollin are correlative with the coastal
The basal quartzitic sandstone and shale of the open plain sandstones observed in cores from the western
marine facies are interpreted to be of subtidal shoal Oriente. Shale interbeds in the coastal plain facies are
origin. These sandstones are generally thinner than the both more numerous and thicker than those observed in
overlying glauconitic shoal deposits and are almost the braidplain deposits.
580 White et al.

Figure 7—Hollin sandstone cross section, Pungarayacu concession, eastern Napo uplift. Inferred correlations demonstrate
the more frequent facies changes in the western Oriente. See Figure 1 for location.

The Upper Hollin deposits present in this roadcut are shows the lithologies and stratigraphic relationships of
part of the shore zone deposition. Individual channels the Napo in the western Oriente basin. The Napo T, U,
have a lenticular geometry 30–150 m or more wide and and M sandstone units are related to a series of regres-
1–3 m thick. The strata are interpreted to be of tidal flat sive-transgressive cycles that built the Napo stratigraphy.
and tidal channel origin. No exposures of the open There were at least four such cycles in the western
marine facies of the Upper Hollin were observed here, Oriente basin. Only the T and U intervals deposited
such as the glauconitic sandstones or carbonates. De sands in the western Oriente basin. To the east in the
Souza Cruz (1989) interpreted the Main Hollin as origi- central Oriente (e.g., Shushufindi field), these sequences
nating from braided fluvial and eolian paleoenviron- are indistinguishable because of their stacked, sand-rich
ments. We see little evidence for eolian deposition. We character which resulted in their amalgamation.
agree with de Souza Cruz that the Upper Hollin in this Mapping of the Napo transgressive shales (referred to as
outcrop is estuarine, although the tidal range need not the lower, middle, and upper Napo shales) define a
have been macrotidal. northeast-southwest Napo shoreline trend within the
western Oriente basin.
Seismic reflection data show that Napo stratigraphy
has substantial acoustic contrasts that can be resolved,
NAPO STRATIGRAPHY AND depending on data quality and signal processing. The
DEPOSITIONAL SYSTEMS most conspicuous acoustic change is at shale-limestone
interfaces. The sandstones generally have gradational
The Napo Formation consists of organic-rich shales, contacts. The least resolvable acoustic contrasts occur
bioclastic grainstones and packstones, and terrigenous within the U and T sandstones where they thin
sandstones believed to have been deposited in fluvial, westward and are difficult to distinguish seismically.
deltaic, marginal marine, and marine shelf environments Geophysical modeling of the U sandstone shows a subtle
during the Late Cretaceous. The Napo Formation amplitude increase where the sandstone is well
conformably overlies the Upper Hollin Formation and is developed. A marked amplitude increase also occurs
in turn overlain unconformably by the Tena Formation where the U sandstone is replaced laterally by limestone.
(Maastrichtian–Paleocene). Total Napo thickness exceeds Seismic models for the T sandstone indicate subtle
275 m over much of the western Oriente basin. Figure 9 amplitude decrease where the sandstone is well
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 581

Figure 8—Roadcut stratigraphy exposed along the Hollin-


Loreto road, eastern Napo uplift. See Figure 1 for location.

developed; however, the overlying B limestone is a


strong reflector that tends to mask the T sandstone. On
this basis, subtle amplitude anomalies in the western
Oriente basin are attributed to the U sandstone.
Amplitude versus offset (AVO) analysis of the U
sandstone indicates that there is no significant offset due
to hydrocarbon-bearing lithologies.

Napo T and U Sandstones


Stacked fluvial and deltaic sandstones comprise the
Napo T and U reservoirs of Shushufindi and Libratador
fields in the central Oriente (Canfield et al., 1982). These
intervals quickly thin and become separated by thicker
marine shales in Sacha and Auca fields (Canfield, 1991).
From these fields westward, both the T and U sand-
stones exhibit different lithofacies (and depositional envi- Figure 9—Composite log of Hollin and Napo formations,
ronments) than in the central Oriente. Quartzose sand- western Oriente basin.
stones occur in each of the western Oriente Napo cores
582 White et al.

bedded sandstones dominate the lower half of the Napo


T and U sandstone packages. In the upper half of the
profile, the sandstone beds are thinner, finer grained,
ripple laminated, and generally glauconitic. Mudstone
interbeds are intermittent within the sandstones.
Capping the T and U intervals are more burrowed
mudstone, minor sandstone (locally thicker and medium
grained), and limestone interbeds. Laminated shales
separate the sandstone intervals from thick carbonate
wackestone and mudstone (B and A limestones, respec-
tively).
Situated between thick intervals of marine limestones
and mudstones, the Napo T and U sandstone packages
show a channel-like development, locally exceeding 10
m thick, that is attributable to fluvial channel, shoreline
estuary, and subtidal shoal origin. Channels and shoals
within this setting were probably controlled by the range
of tidal energy, the shoreline configuration (embayments
and estuaries), and the physiography of the marine shelf
(de Boer et al., 1988; Terwindt, 1988). Westward-flowing
streams delivered sediment to these Napo shorelines.
The initial stacked sandstones of these prograding
channels is inferred to be of fluvial point bar or deltaic
origin.
Overlying the channel sandstones, the beds exhibit
reworking by tidal currents that progressively controlled
sand distribution seaward of the fluvial-dominated
deposition. As in the Upper Hollin, the variety of tidal
environments recognized includes tidal flat, tidal creek,
and subaqueous tidal shoals. Dimensions of these shoals
can range up to hundreds of meters in width, hundreds
of meters to several kilometers in length, and more than
5 m in thickness. Positioning of these sandstone bodies
was probably influenced by paleotopographic highs on
the Napo marine shelf. Glauconitic sandstone shoals are
Figure 10—Idealized Napo depositional package resulting mixed with quartzose shoals in the upper parts of the
from sedimentation following sea level drop on the Napo Napo T and U sequences. The cycle of progradation
marine shelf. dominated by channel and tidal shoal sedimentation was
replaced vertically by mud-dominated marine conditions
examined. Glauconitic sandstones, laminated to followed by transgressive bioclastic and micritic
burrowed mudstones, and fossiliferous limestones are limestone deposition (Figure 11).
common components of both T and U sequences. The
quartzose sandstones exhibit the following characteris-
tics: (1) abrupt basal contacts, (2) bed thicknesses from 30 CRETACEOUS PALEOGEOGRAPHY
cm to 1 m, (3) medium to very fine grain size, (4) large
scale planar and trough cross bedding to ripple lamina- The four principal sandstone packages deposited in
tion, (5) abundant clay drapes along laminations, and (6) the western part of the Oriente basin during
occasional disruption due to burrowing. The glauconitic Aptian–Maastrichtian time were the Main Hollin, Upper
sandstones are similar to those of the Upper Hollin and Hollin, Napo T, and Napo U intervals. The Main Hollin
typically occur in the upper parts of both the T and U is the thickest and most widespread of these intervals. It
successions. Cross beds commonly occur in oblique was deposited initially on an irregular erosional surface.
orientations and less often in strongly oblique, or The valley fill deposits smoothed this relief and created
herringbone, orientation. an alluvial plain that was dominated by braided rivers.
Figure 10 shows a typical vertical profile through the The provenance for the Main Hollin sandstones is
Napo T and U sequences. A varying complex of thin believed to have been the Guyana shield and its
fossiliferous limestones, burrowed silty mudstones, and Paleozoic cover. Grain size decreases from east to west.
thin quartzose sandstones form an interval less than 5 m However, isolated outcrops of Hollin valley fill deposits
thick that commonly occurs above the laminated shales in the uplifts west of Puyo contain gravel- to cobble-sized
below both the T and U. The sandstones abruptly overlie clasts of locally derived igneous basement demonstrating
this complex. Stacked, fine- to medium-grained, cross- the influence of possible local source areas during early
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 583

Figure 11—Napo regressive–transgressive cycle of sedimentation showing rapid progradation followed by sea level rise
with marine onlap.

Hollin sedimentation. The westward extent of the


shoreline during braidplain deposition is unknown
because the Andean fold and thrust belt has now
completely overprinted this area west of the Napo uplift.
The physiography conceptualized for the Main Hollin
braidplain is a very low relief, low gradient system. The
multiple river systems comprising the Hollin alluvial
plain carried a high volume of fine to very coarse sand
that was efficiently distributed over the entire alluvial
plain. The source of the quartzose sand is further
envisaged to have been a much higher relief escarpment
of Paleozoic sandstones and igneous basement. Several
examples of such settings include the valley sandur
deposits of southern Iceland (Bluck, 1974), the Scott
outwash deposits of Alaska (Boothroyd, 1972), the
Canterbury plain of New Zealand (Brown and Wilson,
1988), and the Pantanal escarpment and alluvial plain in
Brazil (Short and Blair, 1986). What these alluvial plains
have in common is a high sand supply and a braided
channel network capable of distributing the high bed
load. Regional to local tectonics and vegetation would
also have had an important impact on the resulting
depositional system. The alluvial plain gradient varies in
these examples from several meters per kilometer to less
than 0.01 m/km in the Pantanal basin.
The Aptian–Maastrichtian eustatic sea level curve
(Figure 12) records significant sea level lows that are
correlated with the regressive-transgressive para-
sequences of the Upper Hollin, Napo T, and Napo U
(Macellari, 1988). The Main Hollin coastal plain and
Upper Hollin shore zone depositional systems are Figure 12—Correlation of Napo stratigraphy with global
believed to have been deposited during the late Albian sea level change for the Early–Late Cretaceous. The T and
sea level rise that inundated the Main Hollin braidplain. U sandstone packages correspond to the significant sea
There is no evidence in the Oriente cores for a significant level lows during the Cenomanian and Turonian, respec-
sea level fall at this time. Deposition of the Upper Hollin tively.
Formation reflects gradual eastward encroachment of the
sea and progressive destruction of the delta plain to form formation of the early Andean orogenic belt. A passive
neritic tidal shoals. Figure 12 shows a more substantial margin shelf apparently received the Hollin and Napo
sea level drop in the Late Turonian at the start of Napo U sedimentation. Limited exposures of phosphatic shales
sedimentation, suggesting a marked westward shift of and cherts in the northwestern Napo uplift suggest the
the shoreline. It is envisaged that Upper Hollin and existence of a shelf slope break and Late Cretaceous
Napo deposition took place on a broad, relatively stable upwelling prior to its destruction during Andean defor-
continental shelf west of the Guyana craton and prior to mation.
584 White et al.

Figure 13—Hollin paleogeography in Albian time. (A) Braided alluvial plain. (B) Initial transgression during Main Hollin
coastal plain deposition. (C) Upper Hollin shore zone deposition in tidally influenced nearshore environments. (D) Open
marine sedimentation ending Hollin sedimentation.

Figures 13 and 14 show a series of block diagrams that Hollin depositional systems. Wells suggest that the
summarize the paleogeography during Hollin and Napo shoreline was close to the Guyana shield at the end of
time. The Albian braided alluvial plain was built on the Hollin deposition. The late Albian maximum flooding
edge of the Guyana shield and covered the Oriente basin event (Lower Napo Shale) essentially closed Hollin sedi-
farther west than the Napo uplift. The position of the mentation.
Albian shoreline has been obliterated by Andean defor- The Napo Formation consists of several transgressive-
mation. Inundation by a late Albian sea level rise estab- regressive packages related to Late Cretaceous eustatic
lished fluvial, deltaic, estuary, and tidal shoal environ- sea level fluctuations (Figure 12) (Haq et al., 1988),
ments (Figures 13b, c, d). The delta and estuarine sand including the Napo T and U (Figure 14). The successive
accumulations now form excellent hydrocarbon reser- parasequences in the Upper Hollin and Napo formations
voirs in addition to the Main Hollin. Sand sedimentation were deposited in a basin with a ramp margin (see Van
rates are inferred to have been very rapid within the Wagoner et al., 1988). This model implies that relative sea
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 585

Figure 14—Napo paleogeography during Cenomanian–Turonian time. (A) Early Cenomanian marine shelf deposition domi-
nating much of the Oriente. (B) Napo T sandstone deposition in the western Oriente. (C) Transgressive marine mud deposi-
tion (Middle Napo Shale). (D) Turonian marine shelf sedimentation prior to sea level lowering and deposition of the Napo U
sandstone cycle.

level did not fall below the shelf break, which precludes RESERVOIR CHARACTERIZATION
lowstand sediments within lowstand fan or prograding
wedge settings. The Hollin and Napo shore zone to shelf Reservoir-Scale Heterogeneity of the Hollin
facies tract transgressed and regressed several times. The Lithofacies
quartzose sandstones of the Upper Hollin, T, U, and M
sequences were deposited after maximum sea level fall The Coca-Payamino and Gacela fields in the north-
and within depocenters (eroded valleys) created during western Oriente basin have sufficient well density to
falling sea level. Eventually both the T and U were allow detailed stratigraphic correlation. Well spacings
inundated and covered with limestone shoals and shelf range up to 3 km in the Gacela field and average about
muds during the subsequent sea level rise. The ramp 1 km along the Coca-Payamino structure. The NNW-SSE
margin model permits major shifts of the shoreline, espe- orientation of the Coca-Payamino is nearly orthogonal to
cially where the rate of sedimentation exceeded the rate the east-west depositional pattern interpreted for the
of subsidence. Hollin strata. Figure 15 is a simplified cross section
586 White et al.

Figure 15—Stratigraphic cross section of the Hollin lithofacies in the Coca-Payamino field, western Oriente basin. Lithofa-
cies have been determined from cored intervals in the field. The top of the Upper Hollin is commonly a succession of thin,
fossiliferous limestones. See Figure 1 for location.

through Coca-Payamino field and illustrates the local intervals. Two sandstone types are present: quartzose
variation of depositional facies interpreted for each well sandstones occur in each facies, while glauconitic sand-
along the structure. Overall, the Coca #4 well contains stones occur only in Upper Hollin and Napo intervals.
the thickest development of the coastal and shelf Figure 17 shows the framework and diagenetic charac-
sequences mainly because of thicker shore zone sand- teristics of the Hollin and Napo sandstones.
stones. A relatively thin veneer of coastal deposits is Quartzose sandstones (Figure 17a) volumetrically
present in each well, except in Coca #7, where the equiv- dominate the arenaceous deposits. Grain size varies
alent interval is dominated by a braided channel. The substantially within a single cored interval. The coarsest
Main Hollin remains consistent throughout the structure. detritus in cores or outcrops occurs in the braidplain
Figure 16 shows the overall lithofacies variations depositional system of the Main Hollin succession. In the
between two wells in the Gacela area immediately south Tiguino #3 core, for example, the braided stream sand-
of the Coca-Payamino field. In the Gacela #1 well, both stones contain beds dominated by coarse to very coarse
the glauconitic sandstones of the shelf and the tidal sand- quartz grains, as well as local quartz granule conglom-
stones of shore zone origin are thicker than their counter- erate lag. The average grain size of the Main Hollin is
parts in the Gacela #2 well. The shelf sandstones in the medium grained. Bimodal grain size segregation in slip-
Gacela #2 are not as glauconite rich as those in Gacela #1 face laminae is typical of much of the cross bedding. The
and have retained significant reservoir porosity. Finally, western Oriente Pungarayacu area has the finest grained
the coastal plain deposits in Gacela #1 appear to be Hollin channel sandstones encountered in the Oriente
absent in Gacela #2. The coastal plain facies is believed to area. Excellent porosity and hydrocarbon staining occur
interfinger with braided channel lithologies. throughout the Oriente in the fine-grained to granule
These two field examples suggest that the Main Hollin textured lithologies. Sandstones in the Upper Hollin and
braided stream sandstones are remarkably consistent in Napo successions also vary significantly in grain size, but
character across each field. The coastal plain, shore zone, generally within the very fine to medium-grained size
and open marine units, by comparison, show significant range; they have locally excellent porosity and perme-
compartmentalization that is largely a function of deposi- ability (Figure 17b).
tional environment. Optimum field development must Glauconitic sandstones of the Upper Hollin shelf
account for this lateral and vertical heterogeneity. facies tract and each of the Napo intervals consist of a
framework of glauconite and quartz grains (Figure 17c).
Sandstone Petrography Glauconite content varies from trace to dominant.
Typically, the glauconite grains are about 200 µm larger
A representative suite of sandstone samples from than associated quartz grains. Whereas the quartz in the
Hollin and Napo facies was examined using standard shelf sand shoal facies was reworked from deltaic and
petrographic techniques. From this analysis, it is shore zone deltas, the glauconite was locally derived by
concluded that similar sandstone framework and diage- diagenetic replacement of biogenic material. Glauconite
netic characteristics occur in each of the reservoir grains are easily compacted under moderate overburden
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 587

Figure 16—Stratigraphic correlation in Gacela field 5 km southwest of Coca-Payamino. Thickness and lithofacies variations
are especially noticeable in the coastal plain and open marine facies.

pressure and may form a pseudomatrix that occludes the The burial diagenetic history of the Cretaceous
original primary porosity. Where the percentage of glau- reservoir sandstones reflects several processes that
conite is less than about 20% of the sandstone occurred in the following order:
framework, the quartz-dominated framework retains
much of the original porosity, resulting in significant • Limited mechanical compaction of framework
reservoir potential. In contrast, the dark green, laterally grains
equivalent glauconitic sandstones are tight due to • Early calcite and pyrite precipitation
framework grain compaction. • Dissolution of unstable framework grains
Quartz dominates the detrital framework in all sand- (feldspars)
stones except the glauconite-rich shelf facies. The quartz • Precipitation of silica overgrowths
is generally monocrystalline and less commonly poly- • Precipitation of kaolinite clay minerals
crystalline; it has a strong undulose extinction. Feldspars
and micas are subordinate to rare, but more abundant in Calcite precipitation occludes the initial porosity in
the Napo T and U sandstones. Feldspar composition thin sandstone beds, especially adjacent to shale
varies from sodic plagioclase to potassic feldspar. Unless interbeds where it forms small, spherulitic concretions.
encased in early calcite cementation, most surviving These calcite-cemented sandstones show no evidence of
feldspar grains exhibit moderate to extensive secondary mechanical compaction, suggesting that protective
leaching. Secondary leaching during burial diagenesis cementation occurred at an early stage. Pyrite precipita-
helped reduce the feldspar content. The provenance is tion in the form of concretionary cements or framboids
believed to be the feldspar-rich granitic Guyana are characteristically associated with the organic debris
basement to the east. However, the possibility of a trapped within the sandstones and shales. Early mechan-
quartzose Paleozoic sandstone source overlying the ical compaction is again limited to isolated grain inter-
basement is also possible. Other components of the penetrations. Organic debris and pyrite crystals are
sandstone include heavy minerals such as zircon and concentrated along stylolite-like surfaces.
coalified plant debris. Silica overgrowths are ubiquitous throughout the
588 White et al.

(a) (b)

(c) (d)

(e) (f)

Figure 17—Petrography of Hollin and Napo reservoir sandstones as seen in thin section photomicrographs and scanning
electron micrographs. (a) Plane light view of Main Hollin braidplain sandstone. (b) Quartzose sandstone of Napo U interval.
(c) Glauconitic sandstone of the Upper Hollin open marine facies. (e) Diagenetic kaolinite occupying isolated pores in thin
section. (f) Scanning electron micrograph of secondary silica overgrowths and kaolinite clay mineral.
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 589

quartz arenites of the Hollin and Napo sequence and this interval. In the quartz-dominated Main Hollin,
provide the framework support that has preserved sediment texture is the primary factor controlling pore
porosity to reservoir depths in the Oriente basin. geometry and connectivity.
Although the overgrowths (Figure 17d) make up only a Figure 18a shows the porosity–permeability data for
small percentage of the sandstones, they strengthen the the entire Upper Hollin in a single Coca-Payamino well.
highly porous sandstones while only slightly reducing Permeability ranges over six orders of magnitude, and
overall primary porosity. Mechanical testing of these no distinct trends are discernible in the overall data set.
sandstones documents the high compressive strength The poorest permeabilities are associated with glau-
required to break the strong silica-cemented framework. conitic sandstones and clay-rich interbeds. Figures 18b
The amount of porosity attributable to framework-grain and 18c illustrate the wide range of measured porosity
dissolution is not significant compared to the primary and permeability in this highly heterogeneous
intergranular porosity preserved by silica overgrowth. formation. Quartz-rich zones are of high reservoir
Precipitation of kaolinite clay minerals followed over- quality and comparable to those found in the Main
growth formation. The kaolinite typically fills small Hollin. Median values for porosity and permeability are
clusters of pores, but does not seriously affect sandstone 8.6% and 1.67 md, respectively. A histogram of grain
permeability (Figure 17e). Kaolinite diagenesis density (Figure 18d) further demonstrates the diversity
succeeded silica overgrowth formation (Figure 17f), but of minerals present in this interval.
preceded oil emplacement. Such relationships are Mercury injection extended range capillary pressure
common and invariably associated with the oil–water data were generated to examine reservoir rock quality,
contact where differentially stained sandstones may determine size and sorting of pore throats, and evaluate
occur below the base of the oil-saturated sandstones. seal capacity. Shales within the Upper Hollin (Figure 19)
are microporous and considered to be effective seals.
Petrophysical Characteristics Because of inhibiting diagenetic effects, glauconitic sand-
stones have bimodal pore throat size distributions and
complex pore geometries (Figure 20). Further reduction
Electric Log Response
in reservoir quality can result from extensive diagenetic
The stratigraphic and sedimentologic characterization pyrite and the abundance of detrital clay drapes and
of the Napo and Hollin reservoirs has been facilitated by coalified plant debris.
using core studies combined with electrical log evalua- Figure 21a shows the porosity-permeability data for
tions. Many of the mineralogic characteristics observed the Main Hollin Formation in a single Coca-Payamino
in cores have a petrophysical log response. Carbona- well. A clear cluster of data in the 15–20% porosity range
ceous debris on cross bed slip faces induces a stronger and greater than 100 md permeability demonstrates
shaly gamma ray response than would be expected from excellent reservoir quality. The Main Hollin is a “clean”
core examination. A clean gamma ray deflection is uniform sandstone, although thin, impermeable clay-rich
typical of a clean sandstone, but a higher gamma interbeds are not uncommon. Figures 21b, c, and d illus-
response may indicate relatively clean sandstones conta- trate the quartz-dominated nature of the Main Hollin.
minated with carbonaceous laminae, shaly sandstone, or Median values for porosity and permeability are 18.6%
carbonaceous limestone or marl. Glauconite and pyrite and 1013 md, respectively. Mercury injection data
reduce the resistivity. The glauconite-rich sandstones (Figure 22) show unimodal well-sorted and well-
result in some of the lowest resistivity responses on connected pores, further substantiating high reservoir
observed logs. Pyrite is locally abundant as a dissemi- quality. Most pore throat radii are larger than 1 µm, with
nated replacement fabric or as concretions in all litholo- most pores greater than 10 µm in width.
gies. Dolomitic shales tend to have higher resistivity than Anisotropy within the Main Hollin causing direc-
nondolomitic shales due to carbonate cementation of tional preferences in permeability is minimal. Horizontal
pore space. These shales are the most resistive clastic and vertical permeabilities were measured on full-
lithofacies in the Oriente basin. diameter core to determine the potential for reservoir
Fluid chemistry is also reflected in log response, and fluid coning. In the quartz-rich zones of the Main Hollin,
its effects limit the usefulness of resistivity or SP curves horizontal and vertical permeabilities are almost equal
for facies correlation. Low salinities within the Main (Figure 23). This indicates that cross bedding and other
Hollin succession limit the reliability of the SP curve and sedimentologic features do not create anisotropy in this
also moderately affect the resistivity curve. The presence sand body.
of oil is noticeable regardless of lithology.
Rock Mechanics
Porosity–Permeability Relationships, Pore
Uniaxial and triaxial compression testing was
Geometry, and Capillarity
performed on four lithologies from the Hollin formation:
Multiple rock types occur in the Hollin formation shale, limestone, glauconitic sandstone, and quartzose
because of variations in depositional environment. The sandstone. These data were critical in the assessment of
most important factors affecting porosity preservation borehole stability and other engineering evaluations
are lithology, compaction, and diagenesis. Porosity and useful for horizontal drilling parameters (Ramirez and
permeability generally correlate in the Upper Hollin Rodas, 1992). Mohr-Coulomb failure criteria were estab-
despite significant mineralogic differences throughout lished under triaxial load on four samples for each
590 White et al.

Figure 18—Porosity–permeability relationships of the Upper Hollin Formation. (a) Porosity versus permeability (to nitrogen
at an estimated net effective reservoir pressure of 2250 psi). (b) Permeability histogram of all lithofacies of the Upper Hollin.
(c) Porosity histogram of all lithologies. (d) Grain density histogram for the Upper Hollin.

lithology. Compressive strengths were measured at based on “fresh state” water-oil relative permeability
9520–25,170 psi for shale, 16,700–28,040 psi for limestone, measurements. Wettability indices in the Upper Hollin
8370–27,550 psi for low-percentage glauconitic support the theory of intermediate to slightly oil-wet
sandstone, and 5100–16,870 psi for clean sandstone conditions. Asphaltinic oils (up to 15.2% asphaltene by
(Figure 24). Tensile strengths ranged from 1760 psi for weight) are common near the oil–water contact. Hollin
shale to 660 psi for clean sandstone. wetting tendencies could have significant impact on
production (fluid flow characteristics) and reservoir
Wettability development scenarios, such as water flood potential.
Both the Upper and Main Hollin demonstrate inter-
mediate to oil-wetting tendencies based on qualitative
and quantitative indicators. Localized development of CONCLUSIONS
mixed wettability or preferentially oil-wet characteristics
can be mineralogy specific, that is, glauconite-rich rocks Core descriptions have shown that four depositional
tend to show stronger oil-wet conditions. Complex pore systems comprise Hollin stratigraphy: braidplain and
geometries formed by small, irregular pore throats lead coastal deposits of the Main Hollin Sandstone, and shore
to high immobile saturation of the wetting phase. Irre- zone and open marine shelf facies in the Upper Hollin
ducible water saturation tends to be low, with an average Formation. This reconstruction enlarges on previous
of 15%, and residual oil saturation ranges from 25 to 40% interpretations of marine-influenced Hollin fluvial depo-
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 591

Figure 19—Petrophysical properties of Upper Hollin shales. (a) Mercury saturation versus injection pressure. (b) Pore size
distribution of reservoir seals in the Upper Hollin.
592 White et al.

Figure 20—Petrophysical properties of Upper Hollin glauconitic sandstones. (a) Mercury saturation versus injection
pressure. (b) Pore size distribution.
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 593

Figure 21—Porosity–permeability relationships of the Main Hollin quartzose sandstones. (a) Porosity versus permeability (to
nitrogen at an estimated net effective reservoir pressure of 2250 psi). (b) Permeability histogram of quartzose sandstone. (c)
Porosity histogram for sandstones. (d) Grain density histogram for the principal reservoir sandstones.

sition. Sandstones in overlying Napo strata in the depositional system of the Upper Hollin succession has
western Oriente basin are also divided into two variable sandstone distribution, with local good quality
sequences (T and U intervals). reservoir development. The capping open marine sand-
The Hollin braidplain depositional system is a stones are moderately prospective, especially where
sandstone-dominated unit that comprises most of the glauconite content is low. Stratigraphic trapping
Hollin succession. It is also the most prolific reservoir potential is implied by the heterogeneity of these litho-
zone in the western part of the basin. The braided fluvial facies.
sandstone units have excellent continuity and connec- Fluviodeltaic Napo sandstones are prolific producers
tivity, as shown by analysis of closely spaced wells. of oil from fields in the central part of the Oriente basin.
However, shale interbeds and thicker channel abandon- These stacked channel and shore zone sandstones have
ment mudstones adversely influence local permeability. reservoir characteristics similar to the underlying Hollin
It is believed that the braidplain deposits are most fluvial sandstone reservoirs, albeit with local hetero-
productive in structural traps where there is limited geneities. Toward the west, the Napo sandstones occupy
stratigraphic trapping potential. valley-like, topographic lows; these sandstones have
The coastal plain depositional system consists of locally significant reservoir potential.
braided and meandering river sediments, overbank A better understanding of the Hollin and Napo
floodplain strata, and deltaic-estuarine deposits. Even stratigraphy and distribution of reservoir quality sand-
between closely spaced wells, sandstone–shale ratios stones will help to optimize wellbore placement during
may be variable. Similarly, the overlying shore zone field development. This understanding has been further
594 White et al.

Figure 22—Petrophysical relationships of the Main Hollin sandstones. (a) Mercury saturation versus injection pressure. (b)
Pore size distribution from mercury injection data.
Reservoir Characterization, Hollin and Napo Formations, Oriente Basin, Ecuador 595

Suarez, and H. J. Welsink, Petroleum basins of South


America: AAPG Memoir 62, this volume.
Bluck, B. J., 1974, Structure and directional properties of some
valley sandur deposits in southern Iceland: Sedimen-
tology, v. 21, p. 533–554.
Boothroyd, J. C., 1972, Coarse-grained sedimentation on a
braided outwash fan, northeast Gulf of Alaska: Coastal
Research Division, University of South Carolina Technical
Report No. 6, 127 p.
Brown, L. J., and D. D. Wilson, 1988, Stratigraphy of the late
Quaternary deposits of the northern Canterbury plains,
New Zealand: New Zealand Journal of Geology, v. 31,
p. 305–335.
Campbell, C. J., 1970, Guide to the Puerto Napo area, eastern
Ecuador, with notes on the regional geology of the Oriente
basin: Ecuador Society of Geology and Geophysics, 40 p.
Canfield, R. W., 1991, Sacha field, Ecuador: Oriente basin, in
N. H. Foster and E. A Beaumont, eds., AAPG Treatise of
Petroleum Geology, Atlas of Oil and Gas Fields, Structural
Traps V, p. 285–305.
Canfield, R. W., G. Bonilla, and R. K. Robbins, 1982, Sacha oil
field of Ecuadorian Oriente: AAPG Bulletin, v. 66,
Figure 23—Full diameter core permeability measurements p. 1076–1090.
(horizontal and vertical) for the Main Hollin sandstones. Cant, D. J., 1982, Fluvial facies models and their application, in
P. A. Scholle and D. Spearing, eds., Sandstone depositional
environments: AAPG Memoir 31, p. 115–137.
Dashwood, M. F., and I. L. Abbotts, 1990, Aspects of the
enhanced by detailed petrophysical analysis of the
petroleum geology of the Oriente basin, Ecuador, in J.
reservoir sandstones, which has provided the appro- Brooks, ed., Classic petroleum provinces: Geological
priate data for accurate reservoir simulation. The Oriente Society of London, Special Publication 50, p. 89–117.
basin of Ecuador is a proven oil province that has de Boer, P. L., A. van Gelder, and S. D. Nio, eds., 1988, Tide-
tremendous potential for future production. influenced sedimentary environments and facies:
Dordrecht, The Netherlands, D. Reidel, 530 p.
de Souza Cruz, C. E., 1989, Cretaceous sedimentary facies and
depositional environments, Oriente basin, Ecuador—a
field trip guide: Tercer Congreso Andino de la Industria
Acknowledgments The authors would like to thank the del Petroleo, Petrobras Research Center, Brazil, 65 p.
Direccion Nacional de Hidrocarburos (DNH) and Petroe- Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic and
cuador for permission to publish this paper and for their Cenozoic chronostratigraphy and cycles of sea-level
invaluable assistance in making Hollin and Napo cores change, in C. K. Wilgus, B. S. Hastings, C. A. Ross, H. Posa-
available. The core examination in Ecuador (Quito and Lago mentier, J. Van Wagoner, and C. G. St. C. Kendall, eds.,
Agrio) was undertaken by the principal author, Ed Robbs and Sea-level change: an integrated approach: SEPM Special
Felix Ramirez (Oryx Energy, Dallas), and Mariana Lascano Publication 42, p. 71–108.
(Petroecuador). Acknowledgment is given for their assistance Macellari, C. E., 1988, Cretaceous paleogeography and deposi-
in collecting the initial core data for the project. Harold Illich tional cycles of western South America: Journal South
American Earth Sciences, v. 1, p. 373–418.
(Oryx Energy, Dallas) contributed substantially to the Hollin
Miall, A. D., 1977, A review of the braided river depositional
outcrop study and our understanding of the Oriente basin environment: Earth-Science Reviews, v. 13, p. 1–62.
burial history. Further acknowledgment is given to Tim Ramirez, F. A., and J. A. Rodas, 1992, Geoscience aspects in
Martin (Oryx Energy, Dallas), Cliff Thomson (Oryx the first experiences with horizontal wells in the Ecuado-
Ecuador), Oryx Energy (Dallas), and our partners for permis- rian Oriente basin: Proceedings, V Congreso Colombiano
sion to publish this paper, and to the Oryx Graphic group for del Petrolero, Memorias, p. 91–100.
preparation of the illustrations. Short, N. M., and R. W. Blair, Jr., 1986, Geomorphology from
space—a global overview of regional landforms: Wash-
ington, D.C., National Aeronautics Space Administration,
715 p.
REFERENCES CITED Smith, D. G., 1987, Meandering river point bar lithofacies
models: modern and ancient examples compared, in F. G.
Almeida, J. P., R. Campania, M. Rivadeneira, F. A. Ramirez, Ethridge, R. M. Flores, and M. D. Harvey, eds., Recent
H. Poveda, H. Gutierrez, C. Cordero, and S. Guevara, 1983, developments in fluvial sedimentology: SEPM Special
El campo de crudos Pesados Pungarayacu: Paper Publication 39, p. 83–91.
presented at the Congreso Ecuatoriano de Geologia, Terwindt, J. H. J., 1988, Palaeo-tidal reconstructions on
Guayaquil, Ecuador. inshore tidal depositional environments, in P. L. de Boer,
Balkwill, H. R., G. Rodrigue, F. I. Paredes, and J. P. Almeida, A. van Gelder, S. D. Nio, eds., Tide-influenced sedimen-
1995, Northern part of Oriente basin, Ecuador: reflection tary environments and facies: Dordrecht, The Netherlands,
seismic expression of structures, in A. J. Tankard, R. D. Reidel, p. 233–263.
596 White et al.

Figure 24—Mohr-Coulomb failure criteria for the Main and Upper Hollin lithologies. (a) Upper Hollin shale. (b) Upper Hollin
limestone. (c) Glauconitic sandstone (high quartz content, Upper Hollin). (d) Main Hollin reservoir sandstone.

Tschopp, H. H. , 1953, Oil explorations in the Oriente of Robert A. Skopec


Ecuador: AAPG Bulletin, v. 27, p. 2303–2347. Department of Petroleum Geology
Van Wagoner, J. C., H. W. Posmentier, R. M. Mitchum, P. R. University of Aberdeen
Vail, J. F. Sarg, T. S. Loutit, and J. Hardenbol, 1988, An Aberdeen AB9 2UE
overview of sequence stratigraphy and key definitions, in
Scotland
C. W. Wilgus, B. S. Hastings, C. A. Ross, H. Posamentier, J.
Van Wagoner, and C. G. St. C. Kendall, eds., Sea-level
changes: an integrated approach: SEPM Special Publica-
tion 42, p. 39–45. Jose A. Rodas
Wasson, T., and J. H. Sinclair, 1927, Geological explorations Oryx Ecuador Energy Company
east of the Andes in Ecuador: AAPG Bulletin, v. 11, Avenue de Amazonas
p. 1253–1281. Quito
Ecuador
Authors’ Mailing Addresses
Howard J. White
Felix A. Ramirez Guido Bonilla
Oryx Energy Company Petroecuador
13155 Noel Road J. Leon M y Av. Orellana
Dallas, Texas 75240-5067 Quito
U.S.A. Ecuador
Intermontane Late Paleogene–Neogene Basins of the
Andes of Ecuador and Peru:
Sedimentologic and Tectonic Characteristics

R. Marocco R. Baudino
ORSTOM Laboratoire de Modélisation des Bassins Sédimentaires
Paris, France Université de Pau
Pau, France
A. Lavenu
ORSTOM
Université de Pau
Pau, France

Abstract

A n important characteristic of Neogene basin evolution in the Andean Cordillera was the formation of
intermontane basins. These basins were initiated in the late Oligocene with reactivation of Andean
tectonism and were abandoned in the latest Miocene (about 7 Ma). Their sedimentary fill and structures record
the Neogene tectonic history. The sedimentary fill of these basins comprises two megasequences. The first
consists of fluvial and overlying lacustrine deposits attributed to basin opening. The second is composed essen-
tially of proximal fluvial sedimentary units and reflects the closure of the basins in the latest Miocene.
Structural analysis of the Neogene basins shows that their evolution was controlled by the regional tectonic
stress. Synsedimentary folding and fracturing show that the direction of stress experienced a clockwise
rotation in the Neogene, thus explaining variations in the behavior of the faults bordering the basins as well as
the different stages of their evolution.

Resumen

U na característica importante de la evolución de las cuencas neógenas de la cordillera de los Andes ha


sido la formación de las cuencas intramontañosas, cuya creación se inició en el Oligoceno superior con
la reactivación de la tectónica andina, y cuyo relleno finalizó en el Mioceno terminal (aproximadamente 7 Ma).
Su relleno sedimentario y sus estructuras registran la historia tectónica del Neógeno. El relleno sedimentario de
tales cuencas comprende dos megasecuencias. La primera consiste en depósitos fluviales y lacustres sobreya-
centes que corresponden a la apertura de las cuencas. La segunda se compone básicamente de sedimentos
fluviales proximales y refleja el cierre de las cuencas en el Mioceno terminal.
Según los resultados del análisis estructural de la cuencas neógenas, su evolución ha sido controlada por el
esfuerzo tectónico regional. Plegamientos y fracturas sinsedimentarios muestran que la dirección del esfuerzo
ha experimentado una rotación horaria durante el Neógeno, explicando así las variaciones en el compor-
tamiento de las fallas del borde de las cuencas y en las diferentes etapas de su evolución.

Marocco, R., A. Lavenu, and R. Baudino, 1995, Intermontane late Paleogene–Neogene 597
basins of the Andes of Ecuador and Peru: sedimentologic and tectonic characteristics, in
A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America:
AAPG Memoir 62, p. 597–613.
598 Marocco et al.

80° 76° 72° 68°

COLOMBIA

lley
an V a
0° Quito 0°

A nde
Coast ECUADOR

Inter-
Guayaquil
Cuenca

M
D -G
4° PERU 4°

Ne
og
Co

en
as
t

e
Fo
Ea

re
s te

la
nd
Cajamarca BRAZIL

rn

Ba
Ma
8° 8°

si
Co
Figure 1—Location of basins in an Andean-type orogenic

in

n
An
rd i
belt.

dea
lle r

nT
AA

hru
st
Hi
SU

gh
BD

lan
UC
INTRODUCTION

ds
TI
12° 12°

ON
Lima

ZO
The phenomena that cause folding and uplift of

NE
Ayacucho Cuzco
orogenic belts are also responsible for the genesis of Al
tip
foreland basins and intermontane basins (Figure 1). The
AAA
We la
s te no
rn
sedimentary fill of these basins and their bounding struc- Nazca
Co

AAA
Co

Pa
as rd i Puno
t lle
tures are controlled by the tectonic forces that build the

ci
16° ra 16°

fic
orogenic belt. Sedimentologic, stratigraphic, and struc-

O
Arequipa
La Paz

ce
tural analysis of the synorogenic basins allows a recon-

an
BOLIVIA

struction of their development and their regional rela-


tionships. This type of study is essential to under- 200 km
standing the geodynamics of mountain belts. 20° 72°
CHILE
80° 76° 68°
We are particularly interested in the Andean inter-
montane basins of Ecuador and Peru because they are
contemporaneous with Neogene tectonism and because Figure 2—Morphostructural map of Ecuador and Peru.
D-G M, Dolores-Guayaquil megashear, the tectonic border
they record structural reactivation. These processes have between the oceanic accreted coastal terranes (on the
significance for the basins described by Mégard et al. west) and the South American continental lithosphere.
(1984), Bonnot et al. (1988), Noblet et al. (1988), Bellier et
al. (1989), Marocco et al. (1990, 1993), and Baudino et al.
(1991). There are three types of intermontane basins: DATING THE BASIN FILL
those linked to strike-slip faulting, those controlled by
reverse faults, and those related to normal faults. Basins The chronostratigraphy of the succession filling the
such as the Andean intermontane basins commonly intermontane Neogene basins is not well known.
develop through each of these types. The term intermon- Sporadic fossil occurrences and radiometric ages date
tane is occasionally misused to define some Andean some specific strata. However, radiometric ages of the
basins. For example, the Moquegua basin of the southern volcanic rocks underlying and overlying the sedimentary
Peruvian coast has been defined as intermontane fill confirm a Neogene age.
(Marocco et al., 1982; Marocco, 1984), although it is
actually a Neogene continental forearc basin. Some small Age of the Volcanic Rocks
basins in the Eastern Cordillera of Perú, which are inter-
preted as Neogene intermontane basins because of their In Ecuador and Peru, the formation of the Neogene
setting, are actually the early stages of the sub-Andean basins began in about the late Oligocene when arc
foreland basin. Examples include the Bagua basin of magmatism recorded the effects of Andean tectonic reac-
northeastern Perú and the Andamarca basin of central tivation. Most recent work on Andean deformation,
Perú. Mathalone and Montoya (1995) illustrate some chronology, and related sedimentation (Noblet et al.,
examples. 1988; Sébrier et al., 1988; Lavenu and Noblet, 1989; Baby
Neogene intermontane basins are located in the inter- et al., 1990; Lavenu et al., 1990; Marocco et al., 1990;
Andean region. This is a region of variable width sepa- Marocco, 1991; Sempere, 1991; Sempere et al., 1991)
rating the Western and Eastern Cordillera of Ecuador agrees that, in the region presently occupied by Neogene
and Perú. It is known as the Inter-Andean Valley in intermontane basins, Andean tectonism was reactivated
Ecuador, as the Highlands (or High Plateau) in central during the late Oligocene. This is the Aymara phase of
Perú, and as the Altiplano in southern Perú and Bolivia Sébrier et al. (1988) and was apparently the start of a long
(Figure 2). tectonic phase (see Sempere, 1991).
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 599

In Ecuador, the Saraguro volcanics immediately 80° 78° 76°

preceding the formation of intermontane basins


COLOMBIA
(Baldock, 1982) have an age bracketed between 35.3 and

rio B
C
26.8 Ma (Lavenu et al., 1992). The oldest dated rocks

lanco
h

within the sedimentary succession yield an age of 22 Ma 0° G 0°


QUITO
(Lavenu et al., 1992); they form the lower part of the

ca
Co
Biblián Formation in the Cuenca basin (Noblet et al.,

ZO NE

r io
*

aule
L
1988). The presence of volcanic units, dated at 19.5–14.2 L

rio D

r io
Ma by Kennerley (1980), is evidence that volcanic activity

SU BD UC TION

Na
po
was persistent during basin subsidence (Marocco et al.,

rio
Riobamba

Cu
GUAYAQUIL R
1993).

ra
2° r io 2°

ra
Ti

y
In Peru, the youngest volcanic series underlying the gr
e

M
C
intermontane basins belongs to the Calipuy Formation in

ain A

rio Pa
the north (Cobbing et al., 1981), to the Huanta, Castrovir- Cuenca

staza
n
reyna, and Ayacucho formations in central Peru G

rio Santia
dean
N
(Mégard, 1978), and to the Tacaza Formation in the south D -G

Thru
M 4°
(Newell, 1949). The Calipuy Formation has been dated at 4°

go
L

54–14 Ma (Cobbing et al., 1981). In central Peru, Noble

st
V 100 km

(1973), Noble et al. (1974), and Dalmayrac et al. (1980)


showed that the volcanic rocks enclosing the intermon- PERU Z

tane basins had an age between 41 and 6 Ma. Finally, the


76°
age of the Tacaza Formation of southern Peru ranges 80° 78°

from 27.2 to 8.9 Ma (Sébrier et al., 1988). Therefore, in


Peru as well as in Ecuador, the enclosing volcanic suites Figure 3—Location of the Ecuadoran Neogene intermon-
were emplaced before and during the evolution of the tane basins. D-G M., Dolores-Guayaquil megashear;
basins. Ch, Chota basin; G, Guayabamba basin; L, Latacunga; R,
The sedimentary fills of the basins are generally Riobamba basin; C, Cuenca basin; G, Girón–Santa Isabel
basin; N, Nabón basin; L, Loja basin; V, Vilcabamba basin;
overlain by volcanic materials. This recent volcanic Z, Zumba basin.
cover, known as the Barroso Group in southern Peru
(Mendivil, 1965), has an age of 7.2–0.17 Ma (Sébrier et al.,
1988). Where present in Ecuador (e.g., Cuenca basin)
(Noblet et al., 1988), the volcanic cover has a younger manco, Poteria [pseudoaperastoma] bibliana), and bivalves
age, from 3.59 Ma to present (Barberi et al., 1988; Lavenu (Ecuadorea bibliana) in the lacustrine intervals of the
et al., 1992). Loyola Formation (Noblet et al., 1988). These indicate a
broadly Miocene age (Bristow and Guevara, 1974).
Notoungulae (Toxodontidae) have been discovered in
Age of Basin Sedimentary Rocks the overlying Mangán Formation (Repetto, 1977). The
ostracod Cyprideis stephensoni and plant remains occur in
Available dates for the sedimentary fill of the inter- the coal at the top of the lacustrine intervals in the Loja
montane basins are both paleontologic and radiometric and Vilcabamba-Malacatos basins (Kennerley and
because significant volcanic activity occurred in the Almeida, 1975; Marocco et al., 1993). These plants,
vicinity of the basins while lacustrine and fluvial studied by Berry (1929), include Camphoromea speciosa,
sediments were being deposited. This is reflected in the Cassia longiflora, Heronymia lehmanni, and Tapirina lanceo-
basins by the occurrence of lava or pyroclastic flows and lata and belong to the late Miocene.
by the deposition of river-transported volcaniclastic Radiometric dates complement the stratigraphic infor-
materials. mation. In some cases, they support the paleontologic
Lacustrine facies present in almost all the basins have ages, such as in the Cuenca basin (Figure 3), where
yielded most of the reported animal and plant fossils, volcanic intercalations show ages of 22–5.2 Ma (early–
which indicate an early–late Miocene age. In the northern late Miocene) (Barberi et al., 1988; Noblet et al., 1988;
Peruvian basins of Namora, San Marcos, and Lavenu et al., 1992), and in the Namora basin of Peru
Cajabamba, Bellier et al. (1989) have described diatomite (Figure 4). Elsewhere radiochronology is the only
associations of early–middle Miocene age in the method for dating the sedimentary column. This is the
Cajabamba Formation and of late Miocene age in the case for the Peruvian Rumichaca basin (see Figure 8),
Namora Formation. In the latter beds, a volcanic tuff where the base of the succession has been dated at 22 Ma
dated radiometrically at 7.2 ± 0.6 Ma confirms the (McKee and Noble, 1982; Mégard et al., 1983) and the
Tortonian age of these fossil associations. Ayacucho basin (see Figure 7), where volcanic intercala-
Lacustrine fauna and flora have been studied in more tions in the volcaniclastic succession yield ages of 18.3–6
detail in Ecuador. In the Cuenca basin, Marshall and Ma (Mégard et al., 1984). In southern Ecuador, the fluvio-
Bowles (1932) and Liddle and Palmer (1941) describe lacustrine sedimentary rocks of the Nabón basin
ostracods (Cyprideis aff. Howei), crabs (Necronectes (Figure 3) have been dated at 22–7.9 Ma (Winkler et al.,
proavitus), gastropods (Limnopomus [ampullarius] cf. 1993).
600 Marocco et al.

80° 76°
COLOM BIA
72° 70°
sequences of the type shown in Figure 9A. The sequence
0° 0° begins with an erosive base and is overlain by a lower
COLOMBIA
conglomeratic or coarse pebbly sandstone bed with
common large-scale trough cross stratification and
pebble imbrication, corresponding to channel fill. Above
ECUADOR
these are finer grained conglomerates of longitudinal bar
4° 4°
origin (see Collinson, 1986), with or without coarse hori-
zontal lamination. The overlying sandstones have hori-
zontal lamination (Figure 9A) or ripple structures attrib-
uted to low water stage. The sequence is commonly
200 km
capped with several meters of claystone or siltstone in

AA
C ajam arca
which small floodplain channels may be incised. These
N BRAZIL

SM 8° sedimentary rocks can be divided into larger decameter-
C ba
or hectometer-scale sequences. The latter are either

AA
C ordillera Blanca batholith
H uaraz fining-upward, as in the lower parts of the sedimentary
fill of the Cuenca (Noblet et al., 1988) and Vilcabamba
M
ai

CH
n

basin (Marocco et al., 1993), or coarsening-upward, as in


An
de

the middle part of the B megasequence of the same


an
Th

basins (Figure 5). The abundance of debris flows interca-


ru

12° Lima
st

12°
A yacucho lated in the fluvial sequences confirms the proximal

AA
R A
C uzco character of the sedimentation.
SU

A cho
BD

P ro Another type of fluvial environment is characterized


U

AA
CT

T Titicaca
by conglomeratic coarsening-upward sequences
IO

Nazca
Pa

Lake
N

10–100 m thick (Figure 9B). These sedimentary units are


cif

ZO

Puno
ic

interpreted as alluvial fans that form the upper part of


Oc

16°
16°
ea

A requipa
the fill in each of the basins studied (Figures 5, 6, 7, 8).
n

La Paz

BOLIVIA These alluvial fans reflect rapid filling of the basins


controlled by tectonic processes that uplifted the basin
CHILE
flanks. The crests of the alluvial fan sequences sometimes
20°
80° 76° 72° 68° contain boulders several meters in diameter, such as in
the upper sequences of the B megasequence in the
Figure 4—Location of the Peruvian Neogene intermontane Vilcabamba basin (Figure 5) (Marocco et al., 1993). The
basins. N, Namora basin; SM San Marcos basin; Cba, conglomerate beds are variable. They may be structure-
Cajabamba basin; CH, Callejón de Huaylas basin; R, less or locally channelized or may consist locally of sheet-
Rumichaca basin; Acho, Ayacucho basin; A, Anta basin; flood deposits, sieve deposits, or debris flows. These
Cc, Ccatca basin; Pro, Paruro basin; T, Tinajani basin. facies types suggest a semi-arid climate during alluvial
fan sedimentation (see Collinson, 1986). The clays that
commonly form the lower parts of the alluvial fan
LITHOSTRATIGRAPHY sequences may contain red oxidation intervals and
The basins are filled with fluvial and lacustrine conti- calcareous crusts, such as in the Vilcabamba basin
nental sedimentary rocks and significant amounts of (Marocco et al., 1993). These argillaceous lower parts
intercalated volcanic material. Because the sedimentary indicate lower energy and reduced coarse sedimentation
environments of the Ecuadorian intermontane basins (such as during tectonic quiescence or long droughts)
have been studied in most detail, we have taken the that favor soil formation.
following examples from the works of Noblet et al.
(1988), Noblet and Marocco (1989), Barragán (1992), and Lacustrine Sedimentary Rocks
Marocco et al. (1993). Lacustrine sediment units are important in the lower
part of the sedimentary fill of each basin. Two main
Depositional Environments types of facies are represented: a sedimentary facies
related to quiescent periods and a catastrophic sedimen-
Only the fluvial and lacustrine sedimentary environ- tary facies contemporaneous with intense tectonic
ments are represented in these intermontane basins activity. The granulometry of the lacustrine fill may thus
(Figures 5, 6, 7, 8). reflect either tectonism or climatic changes. We ascribe
the change from fine-grained to coarse-grained sedimen-
Fluvial Sedimentary Rocks tation to tectonic processes. Although there may be a
The fluvial sedimentary units are typically character- climatic overprint, there are no data available on
ized by fining-upward sequences 0.5–7 m thick (Figure Neogene climates in the Andes.
9A). The relatively small area of Andean intermontane The quiescent sedimentation facies are of two types
basins (rarely exceeding 1000 km2) and the proximity of that reflect their proximity to the shore zone. Clastic sedi-
provenance is reflected in proximal fluvial systems with mentation occurred near river mouths where small
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 601

AAAAAA
AAAA AA A
Chota basin Cuenca basin Vilcabamba basin
1 2 3

AAAA
AA AA
A A AA AA A
1 2 3 Llacao Fm
5000 1 2 3
angular unconformity
2000

AA
AAAAA A A AAA AA A
8 Ma

alluvial fan deposits


AAAAA A A AA AA
AA A
AAAA
AAAAAA A A AA AA A
2000 4000 12 Ma
Mangán Fm 1600

AAA
AA AA
A AA
A AA B

AA AAA B

A
AA
AAA
AAAAAA A AAA AA
A AAA AA
B
16. 3 Ma

AAAAA
A A AA A
AAAAA AA
3000
1200

AAAAAA
AAAA
AAAAAA AA AAAAA
Lacustrine turbidites

AA
AA AAAAA AA
AA
Azogues Fm
1000

AAAA
AAA AA AAAAAA
AAA AA
2000
800
Lacustrine
Loyola Fm turbidites

AAA AA
AA AAA AAAAA
AA AAAA AA
evaporites
local angular unconformity

AAA AA AA AAA AA
20 Ma A
1000 400

AAA A A A AA AAAA AA
A A
Biblián Fm A
A

AAAA
AAA A AA AA
Debris flow

AAAA
Debris flow
0m Om angular unconformity 0m
Paleozoic 26 Ma Loma Blanca Fm
Saraguro Fm

Figure 5—Stratigraphic sections of the Chota (Barragán, 1992), Cuenca (Noblet et al., 1988), and Vilcabamba (Marocco et al.,
1993) basins, Ecuador. Circled letters refer to megasequences A and B. Columns: (1) alluvial fan, (2) fluvial, and (3) lacustrine.

Gilbert-type deltas were formed (Gilbert, 1885) which The catastrophic sedimentary facies are represented
distributed sediments into the lake. The river floodplain by high-density megaturbidites that locally form thick
sediments are mainly argillaceous and suggest swamps successions in the Cuenca (Noblet et al., 1988; Noblet and
where bioturbation and rooting occurred. Biochemical Marocco, 1989) and Girón-Santa Isabel basins (Mediav-
sedimentation may have also occurred in these shallow illa, 1991). Figure 10 shows an idealized lacustrine
zones, resulting in limestone with algae lamination and a megaturbidite sequence containing all the elements
fetid smell due to hydrocarbons, such as in the observed in various outcrops. The lower part of the
Vilcabamba (Figure 5) (Marocco et al., 1993) and sequence consists of debris flows (see Middleton and
Rumichaca basins (Figure 8) (Mégard et al., 1983). Evap- Hampton, 1976; Lowe, 1982) with erosional basal discon-
orites are attributed to lake margin precipitation, such as tinuities that may be channeled but are generally flat
in the Vilcabamba (Figure 5) and San Marcos basins with flute, prod, or groove casts. Clasts in a typical debris
(Figure 6) (Bellier et al., 1989). In deeper offshore zones, flow are centimeter to decimeter scale, whereas intra-
sedimentation was mainly due to suspension processes. clasts of lower lacustrine strata can reach a diameter of
This resulted in shales and very fine grained deposits several meters. Large clasts are commonly concentrated
that are white or pale yellow, well-stratified, and laterally at the top of the bed, indicating that they were trans-
persistent. Slump blocks and slump scars locally indicate ported at the top of the turbidity flow before it solidified
the paleoslope of the lake and suggest some tectonic (Middleton and Hampton, 1976; Lowe, 1982).
instability (e.g., Vilcabamba basin) (Marocco et al., 1993). Sandstones overlying the debris flows are character-
Thin, low-density Bouma-type (1962) turbidite beds of istic of the high-density turbidites described by Lowe
centimeter to decimeter scale are commonly intercalated (1982). The S1 unit (or facies) of Lowe is the lowest. It is
within the fine-grained sedimentary beds. Finally, in all composed of coarse sandstones that are locally microcon-
of the basins, the abundance of interbedded volcanic and glomeratic and preserve some traction structures as well
volcaniclastic strata in the lacustrine fill suggests contem- as poorly developed flat or oblique laminations. The
poraneous volcanic activity. second unit, S2, is composed of thin, centimeter-scale
602 Marocco et al.

AA AA
San Marcos Basin Ayacucho Basin

AA
1 2 3

AA
800
Namora Basin

AA
1 2 3
300

Lacustrine deposits
AA
6 Ma

AA
diatomite

AA
7.2 Ma

AA
200
Ayacucho
Fm

AA
AA
B

AA B

AA AAA
500
7 Ma

AA AAA
7.4 Ma

AA
Om
Cretaceous Molinoyoc volcanics

AAAA
9.0 Ma

AA
angular unconformity
angular unconformity
9.3 Ma

A AA
diatomite

AAA
AA
diatomite

AA
evaporite
A 2000 Upper Huanta

AA
Fm

AA
evaporite
B

AA AA
AAAA AA
AAAA AAAA
0m angular unconformity
Cretaceous

AA
1000
Figure 6—Stratigraphic sections of the San Marcos and

AA
10 Ma
Namora basins, northern Peru. Circled letters refer to
megasequences A and B. Columns: (1) alluvial fan, (2) Lower Huanta
fluvial, and (3) lacustrine. (After Bellier et al., 1989.)

AAAA
Fm

coarsening-upward sandstone beds that formed by


“traction carpet” processes. Their genesis, due to traction
and suspension mechanisms, is linked to the increasing
instability of the turbiditic flow. The S2 facies are rare in
the intermontane basins studied, especially the Cuenca
0m
12 Ma

17. 3 Ma

18. 3 Ma
AA
angular unconformity
Larampuquio
volcanics

basin where lacustrine megaturbidites were recognized.


Noblet and Marocco (1989) attribute this to the low Figure 7—Stratigraphic section of the Ayacucho basin,
south-central Peru. Circled letters refer to megasequences
proportion of spherical grains in the sediment, which is A and B. Columns: (1) alluvial fan, (2) fluvial, and (3) lacus-
composed mainly of idiomorphic volcanic minerals. trine. (After Mégard et al., 1984.)
The S3 facies is the uppermost unit of the high-density
turbidites and consists of medium- to fine-grained
massive sandstones that show dish structures. The S3 Sequence Stratigraphy
unit was deposited by rapid settling of the remaining
sediments still in suspension. Megaturbidite sequences The same pattern of stratigraphic evolution character-
vary from 5 to 20 m in thickness, and their areal extent ized all of the basins. Initial fluvial depositional systems
reaches 250 km2. The S3 facies is commonly capped by a were followed abruptly by lacustrine deposition, and
low density Bouma-type turbidite, either complete fluvial environments were reestablished. The upper
(Ta–Te divisions) or consisting of only the most distal fluvial deposition was progressively more proximal in
components (Tc–Te). In the Cuenca basin, Noblet et al. character and terminated in alluvial fan deposition. The
(1988) have shown that megaturbidites were related to a sedimentary units exhibit a sequential organization that
lacustrine delta, which formed typical coarsening- allows correlation, either among different sections of a
upward mouth bar sequences and received an abundant basin if the sequence-generating processes were local
supply of volcaniclastic material (3000-m level, Cuenca (such as hydrodynamic changes or local morphologic
basin section, Figure 5). modifications) or among basins if the processes were
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 603

AA
Tinajani Basin
AA AA TB S3 S2 S1 DF

AAAA
Quaternary 1 2 3
Rumichaca Basin angular
unconformity

AA AA
low density

AA
1 2 3
600

AA
Upper
Tinajani

AA AA AA
Fm

AA
AA
B

AAAA AAA AA
AA AA AA
AA
1000 angular unconformity

high density
14 Ma

AA
AA
AA
A
AA
AA AA
AAAA
AA

5 - 20 m
Lower

AA
Tinajani

AA
Fm

AAAA A AAAA AA
18 Ma

AAAA
AAAA AA AA
AA
A angular unconformity
19 Ma
Lower

A AA
rhyolitic tuff Tinajani
22 Ma volcanics
27 Ma

AAAA A
basalt 0m angular unconformity

AA

debris flow
0m Puno group
Pucará Group

AA
Figure 8—Stratigraphic sections of the Rumichaca (Mégard
et al., 1983) and Tinajani (Sébrier et al., 1988) basins, south-
central and southern Peru, respectively. Circled letters
refer to megasequences A and B. Columns: (1) alluvial fan,
(2) fluvial, and (3) lacustrine.
Figure 10—Example of a megaturbiditic sequence in the
Azogues Formation, Cuenca basin, Ecuador. TB, Bouma
A B turbidite; DF, debris flow; S1, S2, S3, Lowe’s (1982)
12
0 cm 100 divisions of a megaturbidite. (After Noblet and Marocco,
7 1989.)

AAA
2
(Izquierdo, 1991), Girón-Santa Isabel (Mediavilla, 1991),
and Nabón (Winkler et al., 1993) basins of Ecuador

AAA (Figure 3) and the Cajabamba (Bellier et al., 1989),


Callejón de Huaylas (Bonnot et al., 1988), and Paruro
(Mendivil, 1979) basins of Peru (Figure 4). The
1 Pliocene–Pleistocene Anta and Ccatca basins of the

AAA
0m
Cuzco region in southern Peru (Figure 4) (Cabrera, 1988)
show a fining-upward sedimentary organization, but
conglomerate
clay

coarse
fine
medium

0m they are younger than the basins of this study and are
sandstone
still evolving.
The available chronologic data show that sets A and B
Figure 9—Two types of fluvial sequences observed in the are not the same age in all basins. In the Cuenca basin
sedimentary fill of the Río Chota Neogene intermontane (Noblet et al., 1988) (Figure 5), the age of the disconti-
basins, northern Ecuador. (A) Proximal fluvial sequence, nuity between sets A and B is bracketed between 20 and
lower part of the Río Chota basin sedimentary fill. (B) Two 16.3 Ma. In the Ayacucho basin (Mégard et al., 1984)
sequences (1 and 2) of alluvial fan deposits, upper part of (Figure 7), set A starts at about 15 Ma and finishes at
the Río Chota basin sedimentary fill. (After Barragán, 1992.) about 10 Ma. Finally, in the Tinajani basin, set A was
deposited between 18 and 14 Ma. The term megasequence
used here for convenience to name the sedimentary sets
regional in nature (climatic changes, tectonic modifica- A and B is valid only at the scale of an individual basin.
tions, or eustatic variations in marine basins). This This may reflect the lack of detailed biostratigraphy. The
section addresses the large-scale organization of the fossil flora in the Ecuadorian basins was studied more
deposits or megasequences in the basins. than 50 years ago. A reexamination of these flora may
The intermontane basins shown in Figures 5, 6, 7, and indicate that the megasequences are contemporaneous
8 preserve two types of stratigraphic columns: upward- even between basins.
fining set A and upward-coarsening set B. Thicknesses The discontinuity between megasequences A and B is
vary from hundreds of meters to kilometers in the always well defined. It generally consists of an angular
Cuenca and Vilcabamba (Figure 5), Ayacucho (Figure 7), unconformity, as in the Cuenca (Figure 5), San Marcos
and Tinajani (Figure 8) basins. This two-set organization (Figure 7), and Tinajani basins (Figure 8). However, even
is also observed in other basins, including the Loja without discordances, the discontinuity is expressed by a
604 Marocco et al.

drastic change in the sedimentary rocks, from distal beds Santa Isabel basins are absent. Neither did the
at the top of megasequence A (low-energy lacustrine) to Vilcabamba basin receive as much volcaniclastic material
proximal deposits with a notable increase in grain size at as the Cuenca basin. Turbidites filled the lake and were
the base of megasequence B (lacustrine megaturbidites, overlain by proximal prograding fluvial strata; these in
proximal fluvial). turn were succeeded by typical coarsening-upward
alluvial fan deposits such as those previously described
Megasequence A (Figure 9B). Turbidites or megaturbidites have not been
The lower megasequence (A) begins with proximal recognized in the other basins. Thus, megasequence B
fluvial deposits that grade abruptly into deep lacustrine reflects a significant change in sedimentation from
strata. In the upper part, sequences 20–100 m thick of quiescent lacustrine to progressively coarser. This change
shoaling-upward lacustrine deposits are generally is attributed to the increasing tectonism that affected
present, showing deep facies in their lower parts and basin and provenance areas, as evidenced by the
progressively shallower facies in their upper parts. In the progressive discordances observed in this megasequence
southern Ecuador Vilcabamba basin (Figure 5) (Fierro, in the Cuenca (Noblet et al., 1988), Vilcabamba (Fierro,
1991; Marocco et al., 1993), the top of the lacustrine 1991), and Chota basins (Barragán, 1992).
shoaling-upward sequences preserve swamp facies: In the Namora (Figure 6) (Bellier et al., 1989) and
limestones with algal laminations and a fetid smell and Ayacucho basins (Figure 7) (Mégard et al., 1984), the
coal and evaporite beds that are exploited manually. The coarse alluvial fan deposits forming the top of megase-
thickness of the lower fluvial part of the series varies quence B are overlain by lacustrine deposits of latest
among basins. It exceeds 1000 m in the Cuenca basin, Miocene age, radiometrically dated at 7–6 Ma in the
where it constitutes the Biblián Formation (Figure 5) Ayacucho basin (Ayacucho Formation) and at 7.2 Ma in
(Noblet et al., 1988), whereas it is generally thinner than the Namora basin. This lacustrine sedimentation, which
100 m in the other basins. The Cuenca basin, however, reflects a conspicuous modification of the tectonic regime
has subsided more than the other basins. about 7 Ma, coincides with a discontinuity or interrup-
In summary, during deposition of megasequence A, a tion in the compressive regime that was building the
progressively more distal location of the source areas is central Andean orogenic belt. Such uppermost Miocene
reflected in the general fining-upward trend. This pattern lacustrine strata have not been observed in the other
is interpreted as a sedimentary response to basin subsi- basins, either because they have been eroded or because
dence. The initial depression was controlled by fault volcanic activity concealed their presence. In the Namora
systems, which trended parallel to the latest Oligocene and Ayacucho basins, the local structural framework and
orogenic fabric (roughly north-south in Ecuador and tectonic regime is believed to have resulted in reactiva-
NNW-SSE in Peru) and which concentrated drainage. tion of subsidence, thus creating a lacustrine basin
The earliest deposits of the basin were fluvial. Subse- contemporaneous with lesser amounts of thrusting in
quently, marked subsidence trapped the river systems neighboring areas, which explains the absence of coarse
structurally, resulting in widespread lacustrine deposi- sedimentation.
tion of the lower megasequence in the Neogene Andean
basins. However, the early Miocene was also character-
ized by a eustatic sea level 100 m higher than the present TYPES OF BASIN EVOLUTION
(Haq et al., 1987). If the rate of uplift in the central Andes
during the Neogene (0.15 mm/year) estimated by The tectonic evolution of the Neogene intermontane
Sébrier et al. (1979, 1988) is valid, the lakes must have basins of the central Andes can be grouped into two
been at a relatively low altitude, a few tens or hundreds main types according to the orientation of their
of meters above sea level 20 Ma ago. In this setting, the bounding structures in relation to the orientation of the
high sea level may have maintained a high base level, Neogene stress fields. The first type is the strike-slip
favoring preservation of a lacustrine regime, which was, basin, of which we review the Cuenca basin of Ecuador
nevertheless, controlled by subsidence. as an end-member (Noblet et al., 1988). The second type
of basin is linked to reverse faulting activity; we
Megasequence B emphasize the Rumichaca basin of central Peru (Mégard
Megasequence B is marked in its lower part by the et al., 1983). However, because of the variable nature of
abrupt change from a low-energy sedimentation regime Neogene stress fields, each basin has evolved through
to one of coarse sedimentation. Thus, in the Cuenca basin several stages of the stress regime: extension, transpres-
(Figure 5), quiescent lacustrine sedimentation of the sion, and compression. Each basin is characterized by a
Loyola Formation grades upward into turbiditic and dominant stress stage, the one that lasted longest and
megaturbiditic sedimentation of the Azogues Formation controlled most of the sedimentation of that basin and
(Noblet et al., 1988). This is also characteristic of the deformation of its environs.
Girón-Santa Isabel basin located just southwest of the From a structural perspective, our knowledge varies
Cuenca basin (Mediavilla, 1991) and which was probably greatly from basin to basin, with the Cuenca basin being
connected to it. In the Vilcabamba basin (Figure 5), the the best known. This complicates comparison, especially
base of megasequence B is also composed of conspicuous where an author has emphasized only one aspect of
coarse turbidites (0.5-mm-diameter clasts); however, basin evolution. For example, in the northern Peruvian
megaturbidites such as those in the Cuenca and Girón- basins of Namora, San Marcos, and Cajabamba, Bellier et
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 605

Sedimentation began with establishment of fluvial


systems on either side of the Santa Ana–San Miguel horst
(Figure 12a); drainage of the Biblián Formation was
toward the north and north-northeast. The Santa
Ana–San Miguel horst was probably formed during the
same tectonic events at the end of the Oligocene that
were responsible for creation of the early Cuenca basin.
As the Santa Ana–San Miguel horst disappeared in the
earliest Miocene and depocenters were yoked together
emphasizing the western depocenter, a lacustrine basin
was established. The middle Miocene Loyola Formation
was deposited in this lake (Figure 12b). Near the end of
the middle Miocene, the influx of large volumes of
detrital sediments into the basin (Figure 12c) marked the
beginning of megasequence B. To the south of the basin,
a fluvial system flowed into the lake, forming a delta that
prograded toward the northeast. As a result of sediment
stacking and tectonic instability, the lacustrine delta front
is believed to have collapsed, resulting in the lacustrine
turbidites and megaturbidites of the Azogues Formation.
In the late Miocene, fluvial sediments of the Mangán
Formation prograded into the lake and completely
filled it.
The tectonic evolution of the Cuenca basin can be
interpreted from analysis of the folding and faulting.
Based on a study of the effects of different periods of
synsedimentary folds, Noblet et al. (1988) established a
basin model related to strike-slip movements of the
bordering faults. Figures 13 and 14 show the synsedi-
mentary folds that were formed during three main
compressive periods:

1. The first period coincided with sedimentation of


the Biblián Formation and resulted locally in
conical folds with axes trending N 120° E (Figure
14A) and in sedimentary pinch-outs along faults
trending N 20°–40° E. These structures are
compatible with a dextral strike-slip tectonic
Figure 11—Tectonic setting of the Cuenca basin in Ecuador regime along the approximately north-south
(after Noblet and Marocco, 1989). Key: 1, Mesozoic trending faults with horizontal axes σ1 and σ3
substratum; 2, Saraguro Formation; 3, first megasequence approaching N 30° E and N 120° E, respectively.
(A); 4, second megasequence (B); 5, Andesite of the 2. The second period occurred at the beginning of
Cojitambo dome (6.3 Ma; Barberi et al., 1988).
turbiditic sedimentation of the Azogues Formation.
It was a compressive event that also resulted in
al. (1989) have stressed the extensional phase of basin conical folds (Figure 14B, C) with axes trending
evolution. close to N 150° E and showing a shortening
direction of about N 60° E. These folds observed in
Cuenca Strike-Slip Basin the vicinity of the fault trending N 20°–40° E along
the eastern border of the basin are compatible with
The Cuenca basin is located in southern Ecuador dextral strike-slip movements along that trend.
(Figures 3, 11) and has been studied by Noblet et al. 3. The third prominent compressive period occurred
(1988) and Noblet and Marocco (1989). It is the largest throughout deposition of the Mangán Formation.
Neogene intermontane basin in this part of the Andes Numerous hiatuses affected 2000 m of deposits
that we have studied (100 km × 30 km). It is controlled by (Figure 13). Analysis of these structures and of
faults trending N 170˚ E to north-south and NNE-SSW. those induced by deformation of the underlying
Figure 5 summarizes the overall coarsening-upward beds shows that the fold axes have an approxi-
succession that is more than 4000 m thick in this basin mately north-south orientation (N 170° E in the
and consists of megasequences A and B. Its sedimentary northern part of the basin, N 20°–30° E in the south)
evolution is summarized in Figure 12, where maps (a) and that shortening approaches an east-west
and (b) represent megasequence A and maps (c) and (d) direction. Diagrams D–H in Figure 14 illustrate this
represent megasequence B. third compressional period.
606 Marocco et al.
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 607

Fracture analysis, particularly of the fault systems


bounding the basin (Lavenu and Noblet, 1989), confirms
the model of a basin related to strike-slip movements
interpreted by Noblet et al. (1988) on the basis of fold
deformation. This implies that between the latest
Oligocene and late Miocene–Pliocene phase of basin
subsidence compression directions experienced a
clockwise rotation, from N 20° E to N 60° E (opening
phase of the basin expressed in the Biblián, Loyola, and
Azogues formations) and finally to N 100° E (closing
phase of the basin reflected in the Mangán Formation).
Figure 15 summarizes the geodynamic evolution of the
Cuenca basin, as well as that of the south Ecuadorian
Neogene basins of Nabón, Loja, Vilcabamba-Malacatos,
and Zumba.
The nature of the structural evolution of the Cuenca
basin, the thick sedimentary accumulation comparable to
that of the Vienna pull-apart basin (Royden, 1985), and
the thicker filling in the proximity of the most active
faults all indicate that the Cuenca basin is linked to the
array of strike-slip faults with characteristics similar to
those described by Nilsen and McLaughlin (1985).
The Ecuadorian Vilcabamba (Fierro, 1991; Marocco et
al., 1993) and Girón-Santa Isabel basins (Mediavilla, 1991)
are also of pull-apart type, at least during part of their
evolution (Figure 15). The Neogene basins of Peru are
less well known; some of them may have had a geody-
namic evolution comparable to that of the Cuenca basin.

Rumichaca Reverse Fault Linked Basin


The Rumichaca basin of central Peru (Figure 4) has
been studied by Mégard et al. (1983). It is elongated in a
north-south direction (Figure 16) and has modest dimen-
sions (5 km × 1.5 km). It is limited to the west by a north- Figure 13—Outcrop map and synthetic cross section of the
south trending reverse fault and by the Huapa anticline progressive discordance affecting the Mangán Formation
that deforms the Mesozoic succession. (after Noblet et al., 1988). Legend: 1, Biblián Formation;
2, Cojitambo andesite; 3, Loyola Formation; 4, Azogues
The Rumichaca succession, attributed to the Formation; 5, Mangán Formation; 6, fold axes (third
early–middle Miocene by Petersen et al. (1977), consists synsedimentary tectonic event); 7, probable fault.
of approximately 600 m of continental, partially volcani-
clastic sedimentary rocks (Figure 8). To the east, these
strata rest with angular unconformity on paleorelief in the upper part of this unit has been dated at 22 Ma
carved into the Liassic limestones of the Pucará Group; to (McKee and Noble, 1982). The tuffs are overlain by 100 m
the west, they are in reverse fault contact with Creta- of lacustrine limestones with common algal laminations
ceous sandstones and limestones (Figure 16A). The lower that pinch out toward the west and south. Along the
100 m consists of volcanic tuffs that are variably altered western margin, near the basin-bounding fault, the lime-
and commonly resedimented, with thin decimeter-scale stones contain interbeds of conglomerates with clasts of
intercalations of lacustrine limestone beds. A tuff located Cretaceous sandstones and limestones, indicating fault
activity during the lacustrine calcareous sedimentation.
(facing page) The upper 400 m consists of conglomerates of proximal
Figure 12—Schematic paleogeographic map and synthetic fluvial or alluvial fan origin which reflect conspicuous
cross section for each formation of the Cuenca basin (after activity of the western fault. Clast size of the conglomer-
Noblet et al., 1988). (a) Biblián Formation; (b) Loyola ates increases upward and from east to west.
Formation; (c) Azogues Formation; (d) Mangán Formation. The structure of the basin is dominated by the
Legend: 1, lake; 2, lacustrine megaturbidites; 3, lacustrine Rumichaca syncline (Figure 16). On the Lircay-
delta; 4, flood plain; 5, braided river; 6, alluvial fan; 7, relief; Huachocolpa highway along the Huachocolpa River, the
8, vector transport (N is number of measurements); 9, Rumichaca syncline is markedly asymmetric with a
slump axes; 10, vector transport from some measure-
ments; 11, fault; 12, conical fold axes of second synsedi-
vertical western flank. Detailed mapping of the basin
mentary tectonic event; 13, conical fold axes of third shows that it is bordered to the west by a reverse fault
synsedimentary event. SB1 and SB2 are subbasins of the and that the sedimentary fill of the basin is punctuated
Biblián Formation. by numerous synsedimentary unconformities near the
608 Marocco et al.

reverse fault (“u” in Figure 16). In contrast, the Neogene


strata of the eastern part of the basin are conformable,
demonstrating that the eastern border was tectonically
passive during Neogene sedimentation. From north to
south in the basin, close to the contact with the western
reverse fault, progressively lower levels of the strati-
graphy crop out. The observed structural attitude of the
strata is shown in Figure 16C. This structural–strati-
graphic architecture resembles that along the active
border of the Alto Cardener basin on the southern side of
the Pyrenees (Riba, 1973, 1974).
Synsedimentary deformation of the Rumichaca basin
began after 22 Ma and was related to formation of the
Huapa anticline. Deformation probably ceased at about
10.5 Ma, the age of the Julcani rhyodacitic domes (McKee
and Noble, 1982) that cross cut the structures affecting
the Rumichaca basin fill.
The Río Chota basin in northernmost Ecuador
(Figure 3) has an evolution comparable to that of the
Rumichaca basin during megasequence B deposition.
Marked synsedimentary tectonic activity, with an initial
shortening direction of N 120° E that later changed to
east-west, affected the conglomerates of the upper part of
the succession. These are alluvial fan deposits that
prograded eastward as a consequence of the activity of
the reverse fault system that forms the western margin of
the basin (Barragán, 1992). In contrast, deposition of the
lower megasequence (A) is characterized by a N 130° E
extensional regime. Evolution of the Neogene Nabón
basin of southern Ecuador (Figure 3) is also linked to
activity of a bordering reverse fault (Winkler et al., 1993).
Similar processes controlled the northern Peruvian
Namora, San Marcos, and Cajabamba basins (Bellier et
al., 1989), in which deposition of megasequence A
(Figure 6) was coeval with an extensional regime (σ3
changed from ENE-WSW to northeast-southeast). There,
coarsening-upward megasequence B is coeval with a
compressional regime in which the σ1 directions changed
from ENE-WSW to north-south, with the former normal
faults that controlled deposition of megasequence A
being reactivated as reverse faults.

CONCLUSIONS
Figure 17 summarizes the evolution of the Neogene
intermontane basins of Ecuador and Peru. Their
evolution begun about 28–26 Ma ago with renewed
Andean tectonism after a long period of relative inac-
tivity following the 42-Ma late Eocene compressive
episode (Sébrier et al., 1988). The evolution of these
basins ended at about 7 Ma. The Ayacucho and Tinajani
basins in Peru do not follow the two nearly simultaneous

Figure 14—Schematic structural map of the Cuenca basin


(after Noblet et al., 1988). Legend: 1, Mesozoic substratum;
2, Cenozoic volcanics; 3, Cenozoic sedimentary rocks;
4, travertines; 5, sites of tectonic measurements; 6,
Azogues and Mangán formations; 7, fold axes; 8, fault.
Diagrams AH are explained in text.
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 609

Figure 15—Evolution of the Neogene intermontane basins of southern Ecuador (after Noblet et al., 1988). (A) Location of
probable ancient faults sealed by Saraguro Formation deposits. (B) First synsedimentary tectonic event, with compression
oriented N 30° E. (C) Second synsedimentary tectonic event, with N 60° E compressional direction. (D) Third synsedimentary
tectonic event, with an east-west compressional direction. Dotted pattern is sedimentary deposits synchronous with defor-
mation.

megasequence models; the opening phase in the resulting in opening of the basins. Progressive blocking
Ayacucho ended about 10 Ma (Figure 7) and in the of the wrenching movement caused clockwise rotation of
Tinajani about 14 Ma (Figure 8). There are two possible the stress, until it reached an east-west shortening
explanations for this difference: either basin evolution in direction (about 7 Ma), roughly parallel to the direction
Peru south of Ayacucho is different from that of the of convergence. This caused reverse faulting activity
northern regions or the chronostratigraphy (especially along the north-south faults, which provoked closure of
radiochronology) for all the Peruvian basins needs to be the basins.
revised. The extension observed in the Río Chota basin during
The initiation of basin formation and subsidence deposition of megasequence A is difficult to interpret. It
coincided with the last major reorganization of the is perpendicular to the compression direction that
Pacific oceanic plates at 27–25 Ma, which caused affected the more southward regions (such as the Cuenca
partition of the Farallon plate into the Cocos and Nazca and Vilcabamba basins) during the same period. Because
plates (Handschumacher, 1976; Pilger, 1984). This new of local characteristics, extensional structures in the Río
organization was associated with reorientation (N 80° E) Chota basin probably presented a better expression than
and acceleration of the convergence rate between the compressive structures. The similarities in patterns of
Nazca and South American plates (Minster and Jordan, evolution of the Ecuadorian basins is also observed in the
1978; Pilger, 1983; Duncan and Hargraves, 1984). Girón–Santa Isabel (Mediavilla, 1991), Nabón (Winkler et
Taking global geodynamic reorganization into al., 1993), and Loja (Izquierdo, 1991) basins, including the
account, Lavenu and Noblet (1989) proposed a model for ages of the sequences and their bounding surfaces and
the evolution of the Ecuadorian Neogene basins (Figure their tectonic characteristics. Only the Zumba basin,
14). The 50° obliquity between the convergence direction straddling the Ecuador–Peru border, is largely unknown.
and the orientation of the Ecuadorian active margin The Peruvian margin, at least south of the
caused the northward displacement of the coastal block, Huacabamba deflection (south of 5° S lat), does not
which accreted in the latest Cretaceous (Mégard et al., comprise accreted terranes, which may explain the
1986). This displacement induced a dextral translation important differences in geodynamic evolution seen
movement along the preexisting north-south faults, among the Neogene basins of Peru and Ecuador. Basin
610 Marocco et al.

Figure 16—Geologic map and cross sections (AC) of the Rumichaca basin (after Mégard et al., 1983). J, Jurassic (Pucara
Group); Ki, Lower Cretaceous (Goyllarisquizga Formation); Ks, Upper Cretaceous (Chulec Formation); 1, volcanics; 2, lacus-
trine limestones; u, intraformational angular unconformity.

evolution in Peru has been controlled by active margin or compression that was oblique to the preexisting
processes and Andean fold and thrust belt deformation faults controlling the basins. This opening phase of
since the late Oligocene. Simultaneous with eastward basin formation lasted until the end of the early
encroachment of the thrust belt, the Andes were uplifted. Miocene, that is, when the main stress directions
The adaptation of preexisting structures to these two changed from NNE-SSW to northeast-southwest
processes explains the origin of the Neogene intermon- (Figure 15).
tane basins. Bellier et al. (1989) interpreted the extension 2. The second stage represented by megasequence B
during the deposition of megasequence A in the defines basin closure during the middle–late
Namora, Cajabamba, and San Marcos basins as a result Miocene as compression progressively approached
of the flow of Andean material toward the trench due to an east-west orientation. Movement along the
gravity forces and weak coupling within the subduction bordering faults in a reverse sense caused uplift of
zone. If this explanation is correct, then the presence in the margins and influx of progressively coarser
Ecuador of a coastal terrane between the trench and the sediments that overfilled the basin.
rising Andes must have opposed this gravity flow.
Evolution of the Neogene intermontane basins of the There is insufficient stratigraphic information to
Andes of Peru and Ecuador follows two stages: correlate megasequences A and B of the Neogene inter-
montane basins with those of the coeval sub-Andean
1. The first stage represented by megasequence A foreland basin or coastal basins with confidence. In the
spans the establishment of the initial basin in which sub-Andean basin of northern Peru, Marocco (1993) has
progressively more distal sediments accumulated. shown that sedimentation proceeded in three coars-
This period was contemporaneous with extension ening-upward megasequences (sequence N1, 28–10 Ma;
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 611

ECUADOR PERU
central Andes at 28–26, 17–15, 10, 7, and 2.7 Ma (Sébrier
et al., 1988)? The angular discordances possibly express
Namora
Vilca- Ayacucho Rumi- Tinajani short breaks or kinematic modifications in the tectonic

AAA
AAA
Chota Cuenca S. Marcos chaca
bamba Cajabamba
continuum. Sempere (1991) has come to a similar conclu-

AAAAAA
ca 7 Ma
sion in his study of the Cenozoic basins of Bolivia.
B

AAA
Upper
Miocene
B
11 Ma
B B B B

Middle
Miocene A B Acknowledgments We thank T. Sempere and an
anonymous reviewer for their discussions which have helped us

AAA
? to improve the original manuscript. We are also grateful to J.
16 Ma

AAA AAAA
? ? ? A
Delfaud and E. Jaillard for helpful discussions. This work has

AAA AAAA
Lower been supported by ORSTOM and IFEA.
Miocene A

AAA
AAAAAAA
A A
A

23.5 Ma
AAAA REFERENCES CITED
AAAAAAA
A ?
Upper

AAAA
AAA AAAAAAA
Oligocene
26-28 Ma Baby, P., T. Sempere, J. Oller, L. Barrios, G. Hérail, and R.

AAAA
AAA
BASIN Paleogene Lower
SUBSTRAT. Paleozoic volcanics
Cretaceous Mesozoic Liassic Oligocene Marocco, 1990, Un bassin en compression d’âge oligo-

AAA
miocène dans le sud de l’Altiplano bolivien: Comptes
Cenozoic volcanism Extension Compression
Rendus Académie des Sciences de Paris, t. 311, sér. II, p.
341–347.
Baldock, J. M., 1982, Geología del Ecuador, boletín explicativo
Figure 17—Summary of the evolution of the Andean del mapa geológico del Ecuador, escala 1/1,000,000:
Neogene intramontane basins. Dirección General de Geología y Minas, Quito, Ecuador,
66 p.
Barberi, F., M. Coltelli, G. Ferrara, F. Innocenti, J. M. Navarro,
sequence N2, 10–7 Ma; and sequence N3, 7–2.7 Ma), thus and R. Santacroce, 1988, Plio-Quaternary volcanism in
forming a coarsening-upward succession. Sequences N1, Ecuador: Geological Magazine, v. 125, p. 1–14.
N2, and N3 represent three stages in the eastward propa- Barragán, R., 1992, Evolución geodinámica de la cuenca
gation of the Main Andean thrust (Figure 2). Although terciaria del Río Chota, provincia de Imbabura: Ph.D.
tenuous, we correlate the intermontane megasequences dissertation, Engineering Dept., Escuela Politécnica
A and B with the first two sequences (N1 and N2) of the Nacional, Quito, Ecuador, 149 p.
sub-Andean foreland basin. Baudino, R., R. Barragán, and R. Marocco, 1991, Nuevos datos
sobre la estratigrafía de la cuenca del Rio Chota (abs.):
West of the continent, the few available syntheses of Sexto Congreso Ecuatoriano Ingeniería Geología Minas y
the Ecuadorian coastal forearc marine basins of Neogene Petróleo, Guayaquil, Ecuador, p. 84.
age (Baldock, 1982; Egüez et al., 1991) show that their fills Bellier, O., M. Sébrier, F. Gasse, E. Fourtanier, and I. Robles,
are organized into four fining-upward sequences: N1 1989, Évolution géodynamique mio-pliocène et quater-
(28–26 to 17–15 Ma), N2 (17–15 to 10 Ma), N3 (10–7 Ma), naire des bassins de la cordillère occidentale du Nord-
and N4 (7–2 Ma). The geodynamics of the coastal basins Pérou: les bassins de Cajabamba, San Marcos et Namora
are controlled by prominent eustatic variations and indi- (Département de Cajamarca): Géodynamique, Paris, v. 4,
rectly by Andean tectonics, which in each pulse supplied p. 93–118.
coarser detrital deposits to the coastal basins, thus Berry, E. W., 1929, The fossil flora of the Loja basin in
forming the base of the megasequences. It is impossible southern Ecuador: Johns Hopkins Studies in Geology,
Baltimore, v. 10, p. 79–125.
to establish intrabasin chronologic correlations of Bonnot, D., M. Sébrier, and J. Mercier, 1988, Évolution géody-
sequences in the three morphostructural regions (coast, namique plio-quaternaire du bassin intra-cordillérain du
Andean zone, and sub-Andean zone) because the Callejón de Huaylas et de la Cordillère Blanche, Pérou:
processes controlling the evolution of the basins in each Géodynamique, Paris, v. 3, p. 57–83.
region were different. Bouma, A. H., 1962, Sedimentology of some flysch deposits: a
The results of the study of the intermontane basins of graphic approach to facies interpretation: Amsterdam,
Ecuador and Peru expand our understanding of Elsevier, 168 p.
Neogene tectonism in both countries (Dalmayrac et al., Bristow, C. R., and S. Guevara, 1974, Hoja geológica al
1980; Sébrier et al., 1988). During the Neogene, the Andes 1/50,000 de Gualaceo: Servicio Nacional de Geología y
experienced persistent compression, as demonstrated by Minería, Quito, Ecuador.
Cabrera, J., 1988, Néotectonique et sismotectonique dans la
the synsedimentary discordances and the chronology of Cordillère Andine au niveau du changement de géométrie
brittle deformation observed in most of the basins. For de la subduction: la région de Cuzco (Pérou): D.Sc. disser-
the Neogene at least, the hypothesis of deformation by tation, University of Paris–Sud, Orsay, 275 p.
short tectonic phases separated by long periods of Cobbing, E. J., W. S. Pitcher, J. J. Wilson, J. M. Baldock, W. P.
tectonic quiescence is incorrect. What, then, is the signifi- Taylor, W. McCourt, and N. J. Snelling, 1981, The geology
cance of the main tectonic “phases” marked by angular of the Western Cordillera of northern Peru: Institute of
unconformities that occurred simultaneously in the Geological Sciences, London, Overseas Memoir 5, 143 p.
612 Marocco et al.

Collinson, J. D., 1986, Alluvial sediments, in H. G. Reading, Marocco, R., 1991, Sedimentación neógena continental en los
ed., Sedimentary environments and facies: Oxford, Andes Centrales, implicaciones geodinámicas: Sexto
Blackwell Scientific Publications, 615 p. Congreso Geológico Chileno, Actas 1, p. 690–693.
Dalmayrac, B., G. Laubacher, and R. Marocco, 1980, Géologie Marocco, R., 1993, Sedimentación neógena en el nororiente
des Andes Péruviennes: Travaux et Documents de peruano, implicancias geodinámicas: First International
l’ORSTOM, 501 p. Seminar: Improvements in Practices of Oil and Gas Explo-
Duncan, R. A., and R. B. Hargraves, 1984, Plate tectonic ration, November 1993, Lima, v. 2, p. 1–24.
evolution of the Caribbean region in the mantle reference Marocco, R., J. Delfaud, F. Mégard, and M. Sébrier, 1982, Une
frame, in W. E. Bonini, R. B. Hargraves, and R. Shagam, série continentale d’un bassin intramontagneux des Andes
eds., The Caribbean–South American plate boundary and Centrales: le Groupe Moquegua (Sud du Pérou): Abstract
regional tectonics: Geological Society of America Memoir 9e Réunion Annuelle des Sciences de la Terre, Société
162, p. 81–93. Géologique de France, p. 178.
Egüez, A., R. Marocco, and V. H. Perez, 1991, Memoria Marocco, R., A. Lavenu, and C. Noblet, 1990, La cuenca intra-
técnica del mapa tectónico del Ecuador: Informe inédito montaña en compresión de Vilcabamba (Sur del Ecuador):
del proyecto EPN-CLIRSEN-IPGH-ORSTOM, Quito análisis tecto-sedimentario (extended abs.): International
(unpublished report), 1 map, scale 1/1,000,000, 44 p. Symposium on Andean Geodynamics, Grenoble,
Fierro, J., 1991, Evolución geodinámica neógena de la cuenca p. 285–288.
intramontañosa de Malacatos-Vicabamba: Ph.D. disserta- Marocco, R., A Lavenu, and J. Fierro, 1993, Sedimentación
tion, Engineering Dept., Escuela Politécnica Nacional, continental neógena en contexto tectónico: la cuenca de
Quito, Ecuador, 149 p. Vilcabamba-Malacatos (Sur del Ecuador): Boletín
Gilbert, G. K., 1885, The topographic features of lake shores: Geológico del Ecuador, v. 3, p. 1–28.
USGS Annual Report, v. 5, p. 75–123. Marshall, W. B., and E. A. Bowles, 1932, New fossil fresh-
Handshumacher, D. W., 1976, Post-Eocene plate tectonics of water mollusks from Ecuador: Proceedings, U.S. Natural
the eastern Pacific, in G. H. Sutton et al., eds., The History Museum, v. 82, p. 1–7.
geophysics of the Pacific Ocean basin and its margin: Mathalone, J., and M. Montoya, 1995, Petroleum geology of
American Geophysical Union, Geophysical Monograph 19, the sub-Andean basins of Peru, in A. J. Tankard, R. Suarez,
p. 177–202. and H. J. Welsink, Petroleum basins of South America:
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of AAPG Memoir 62, this volume.
fluctuating sea levels since the Triassic: Science, v. 235, McKee, E. H., and D. C. Noble, 1982, Miocene volcanism and
p. 1156–1167. deformation in the Western Cordillera and high plateaus
Izquierdo, O., 1991, Estudio geodinámico de la cuenca intra- of south-central Peru: GSA Bulletin, v. 93, p. 657–662.
montañosa cenozóica de Loja (Sur del Ecuador): Ph.D. Mediavilla, J., 1991, Evolución geodinámica de la cuenca
dissertation, Engineering Dept., Escuela Politécnica terciaria de Girón–Santa Isabel, Sur del Ecuador: Ph.D.
Nacional, Quito, Ecuador, 139 p. dissertation. Engineering Dept., Escuela Politécnica
Kennerley, J. B., 1980, Outline of the geology of Ecuador: Nacional, Quito, Ecuador, 210 p.
Institute of Geological Sciences (London), Overseas Mégard, F., 1978, Étude géologique des Andes du Pérou
Geology and Mineral Resources, v. 55, 17 p. Central: Mémoires ORSTOM, Paris, v. 86, 310 p.
Kennerley, J. B., and L. Almeida, 1975, Hojas geológicas de Mégard, F., R. Marocco, J. C. Vicente, C. Muñoz, R. Pastor,
Loja y Gonzanamá: Servicio Nacional de Geología y and J. Mégard-Galli, 1983, Apuntes sobre la geología de
Minería, Quito, Ecuador. Lircay (Huancavelica-Perú Central). El plegamiento tardi-
Lavenu, A., and C. Noblet, 1989, Synsedimentary tectonic hercínico y las modalidades del plegamiento andino (fase
control of Andean intermontane strike-slip basins of south Quechua): Boletín Sociedad Geológica del Perú, v. 71,
Ecuador (South America): International Symposium on p. 255–262.
Intermontane Basins, Geology and Resources, Chiang Mai, Mégard, F., D. C. Noble, E. H. McKee, and H. Bellon, 1984,
Thailand, p. 306–317. Multiple pulses of Neogene compressive deformation in
Lavenu, A., C. Noblet, and T. Winter, 1990, Neogene stress the Ayacucho intermontane basin, Andes of central Peru:
pattern in southern Ecuador (extended abs.): International GSA Bulletin, v. 95, p. 1108–1117.
Symposium on Andean Geodynamics, Grenoble, Mégard, F., M. Lebrat, and T. Mourier, 1986, Las suturas entre
p. 211–214. bloques exóticos y continente en el Ecuador y el norte del
Lavenu, A., C. Noblet, M. G. Bonhomme, A. Egüez, F. Dugas, Perú: Comunicaciones, Santiago de Chile, v. 37, p. 17–30.
and G. Vivier, 1992, New K-Ar age dates of Neogene and Mendivil, S., 1965, Geología de los cuadrángulos de Maure y
Quaternary volcanic rocks from the Ecuadorian Andes: Antajave: Boletín Comisión Carta Geológica Nacional,
implications for the relationships between sedimentation, Lima, v. 10, p. 125.
volcanism, and tectonics: Journal of South American Earth Mendivil, S., 1979, Estratigrafiá de la fase tardigeoliminar en
Sciences, v. 5, p. 309–320. el Perú meridional: Boletín Sociedad Geológica del Peru, v.
Liddle, R. A., and K. M. V. Palmer, 1941, The geology and 60, p. 267–283.
paleontology of the Cuenca-Azogues-Biblián region, Middleton, G. V., and M. A. Hampton, 1976, Subaqueous
Provinces of Cañar and Azuay, Ecuador: American sediment transport and deposition by sediment gravity
Bulletin of Paleontology, v. 26, p. 360–421. flows, in D. J. Stanley and D. J. P. Swift, eds., Marine
Lowe, D. R., 1982, Sediment gravity flows II: depositional sediment transport and environmental management: New
models with special reference to the deposits of high- York, John Wiley, p. 197–218.
density turbidity currents: Journal of Sedimentary Minster, J. B., and T. H. Jordan, 1978, Present-day plate
Petrology, v. 52, p. 279–297. motion: Journal of Geophysical Research, v. 83 (B11),
Marocco, R., 1984, Dynamique du remplissage d’un bassin p. 5331–5354.
intramontagneux cénozoïque andin: le Bassin Moquegua Newell, N. D., 1949, Geology of the Lake Titicaca region, Peru
(Sud du Pérou): Cahiers ORSTOM, série Géologie, Paris, v. and Bolivia: GSA Memoir 36, 111 p.
14, p. 117–140.
Intermontane Late Paleozoic–Neogene Basins of the Andes of Ecuador and Peru 613

Nilsen, T. H., and R. J. McLaughlin, 1985, Comparison of Sébrier M., A. Lavenu, M. Fornari, and J. P. Soulas, 1988,
tectonic framework and depositional patterns of the Tectonics and uplift in central Andes (Peru, Bolivia, and
Hornelen strike-slip basin of Norway and the Ridge and northern Chile) from Eocene to present: Géodynamique,
Little Sulphur Creek strike-slip basins of California, in K. T. Paris, v. 3, p. 139–161.
Biddle and N. Christie-Blick, eds., Strike-slip deformation, Sempere, T., 1991, Cenozoic tectonic “phases” in Bolivia: some
basin formation, and sedimentation: SEPM Special Publi- needed clarifications (abs.): Sexto Congreso Geológico
cation 37, p. 79–103. Chileno, Actas 1, p. 877–881.
Noble, D. C., 1973, Tertiary pyroclastic rocks of the Peruvian Sempere, T., G. Hérail, J. Oller, and M. G. Bonhomme, 1990,
Andes and their relation to lava volcanism, batholith Late Oligocene–early Miocene major tectonic crisis and
emplacement, and regional tectonism: GSA Abstracts with related basins in Bolivia: Geology, v. 18, p. 946–949.
Program, v. 5, p. 86–87. Sempere, T., G. Hérail, P. Baby, R. Marocco, J. Oller, and L.
Noble, D. C., E. H. McKee, E. Farrar, and U. Petersen, 1974, Barrios, 1991, El Altiplano boliviano—una provincia de
Episodic Cenozoic volcanism and tectonism in the Andes cuencas intramontañas de antepaís relacionadas con el
of Peru: Earth and Planetary Science Letters, v. 21, acortamiento cortical en la región del oroclino boliviano:
p. 213–220. Revista Técnica YPFB, Santa Cruz, v. 12, p. 225–227.
Noblet, C., and R. Marocco, 1989, Lacustrine megaturbidites Winkler, W., A. Egüez, D. Seward, M. Ford, F. Heller, D.
in an intramontane strike-slip basin: the Miocene Cuenca Hungerbühler, and M. Steinmann, 1993, A short-lived
basin of south Ecuador: International Symposium on Inter- compression-related sediment fill in the Andean intermon-
montane Basins: Geology and Resources, Chiang Mai, tane basin of Nabón (late Miocene, southern Ecuador)
Thailand, p. 282–293. (extended abs.): Second International Symposium of
Noblet, C., A. Lavenu, and F. Schneider, 1988, Étude géody- Andean Geodynamics, Oxford, England, ORSTOM, Paris,
namique d’un bassin intramontagneux tertiaire sur p. 321–324.
décrochements dans les Andes du Sud de l’Équateur:
l’exemple du bassin de Cuenca: Géodynamique, Paris, v. 3,
p. 117–138.
Petersen, U., D. C. Noble, M. J. Arend, and P. G. Goodel, 1977,
Geology of the Julcani mining districts, Peru: Economic
Geology, v. 72, p. 931–949.
Pilger, R. H., 1983, Kinematics of the South American subduc-
tion zone from global plate reconstructions: Geodynamics
Series, v. 9, p. 113–125. Authors’ Mailing Addresses
Pilger, R. H., 1984, Cenozoic plate kinematics, subduction and R. Marocco
magmatism: South American Andes: Journal of the Fozieres
Geological Society of London, v. 141, p. 793–802. 39700 Lodeve
Repetto, F., 1977, Un Mamífero fosil nuevo en el terciario del
Ecuador (Azuay-Cañar): Tecnología (Guayaquil), v. 1,
France
p. 33–38.
Riba, O., 1973, Las discordancias sintectónicas del Alto A. Lavenu
Cardener (Prepireneo catalán), ensayo de interpretación ORSTOM
evolutiva: Acta Geologica Hispanica, v. 8, p. 90–99. Université de Pau
Riba, O., 1974, Tectongénèse et sédimentation: deux modèles Avenue de l’Université
de discordances syntectoniques pyrénéennnes: in Associa- Pau
tion des Géologues du Sud Ouest, Journées de Toulouse, France
p. 85–103.
Royden, L. H., 1985, The Vienna basin: a thin-skinned pull- R. Baudino
apart basin, in K. T. Biddle and N. Christie-Blick, eds.,
Strike-slip deformation, basin formation, and sedimenta-
Laboratoire de Modélisation des Bassins
tion: SEPM Special Publication 37, p. 319–338. Sédimentaires
Sébrier, M., R. Marocco, J. J. Gross, S. Macedo, and M. Université de Pau
Montoya, 1979, Evolucíon neógena del Piedemonte Avenue de l’ Université
pacífico de los Andes del Sur de Perú: Segundo Congreso Pau
Geológico Chileno, actas 3, p. 171–188. France
Basin Development in an Accretionary, Oceanic-
Floored Fore-Arc Setting: Southern Coastal Ecuador
During Late Cretaceous–Late Eocene Time

Étienne Jaillard Martha Ordoñez


ORSTOM Stalin Benitez
Paris, France Gerardo Berrones
Nelson Jiménez
Galo Montenegro
Italo Zambrano
Petroproducción
Guayaquil, Ecuador

Abstract

S outhern coastal Ecuador is an accreted terrane underlain by an oceanic crust formed during the Aptian-
Albian. To the southeast, the oceanic crust is overlain by Cenomanian–Coniacian fine-grained pelagic
deposits, coarse-grained volcaniclastic turbidites of Santonian–Campanian age, and Maastrichtian–middle
Paleocene tuffaceous shales. Toward the northwest, late Campanian–Paleocene volcaniclastic beds and lava
flows of island arc composition rest on the oceanic crust. This results from the opening of a marginal basin
between an early Late Cretaceous island arc (Cayo arc) and a latest Cretaceous–Paleocene island arc (San
Lorenzo arc).
In the late Paleocene, the accretion of the Cayo remnant arc to the Andean continental margin caused a major
deformation phase that affected only the southern part of coastal Ecuador. There, deformation was sealed by
thick, coarse-grained, quartz-rich turbidites that constitute the infilling of an early fore-arc or slope basin. A
subsequent tectonic event in the early Eocene is believed to have resulted in emergence of the entire area.
At the early–middle Eocene boundary, new fore-arc basins were created that filled with mud and clastic
shelf deposits. A marked disconformity is overlain by coastal to continental coarse-grained deposits of late
middle–early late Eocene age. These express a major tectonic phase attributed to definitive collision of coastal
Ecuador with the Andean margin. The entire area then emerged, until the formation of new fore-arc basins in
the latest Oligocene–Miocene.
The late Paleocene, earliest Eocene, and early late Eocene tectonic events are the most important deforma-
tion phases to affect southern coastal Ecuador and represent its progressive accretion to the margin. The
creation of repeated fore-arc basins can be attributed to subsidence from crustal erosion of the upper plate
because each subsidence event succeeded an important compressive phase that must have favored coupling
and tectonic erosion. This complex geologic history has implications for burial and maturation of organic
matter and must be taken into account in guiding oil exploration in coastal Ecuador.

Resumen

L a Costa Sur del Ecuador es un terreno acrecionado formado por una corteza oceánica que se formó en el
Aptiano-Albiano. Al Sureste, fué cubierta por depósitos pelágicos finos de edad Cenomaniano-Conia-
ciano, seguidos por turbiditas volcanoclásticas gruesas del Santoniano-Campaniano y por lutitas tobaceas de
edad Maastrichtiano-Paleoceno medio. Al Noroeste, turbiditas volcanoclásticas gruesas y coladas volcánicas
de arco insular, datadas del Campaniano-Paleoceno descansan sobre la corteza oceánica. Estos sedimentos se
depositaron en una cuenca marginal que se abrió entre un arco insular activo durante la parte temprana del
Cretácico superior (arco Cayo) y un arco insular, activo en el Cretácico terminal y Paleoceno (arco San
Lorenzo).

Jaillard, É., M. Ordoñez, S. Benitez, G. Berrones, N. Jiménez, G. Montenegro, and I. Zambrano, 615
1995, Basin development in an accretionary, oceanic-floored fore-arc setting: southern coastal
Ecuador during Late Cretaceous–Late Eocene time, in A. J. Tankard, R. Suárez S., and H. J.
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 615–631.
616 Jaillard et al.

En el Paleoceno superior, una fase de deformación mayor que afectó solo la parte Sur de la costa ecuatoriana
representa probablemente la colisión del arco remanente Cayo contra la margen andina. Está sellada por
potentes turbiditas gruesas ricas en cuarzo que constituyen el relleno de una primera cuenca de ante-arco o de
talud. Un nuevo evento tectónico importante en el Eoceno inferior provocó probablemente la emersión de todo
el area.
En el límite Eoceno inferior-medio, una segunda cuenca de antearco se formó y fué rellenada por sedi-
mentos lutáceos y arenosos de plataforma. Una discontinuidad está cubierta por depósitos gruesos costeros o
continentales datados del fin del Eoceno medio y base del Eoceno superior. Estos depósitos expresan una fase
mayor relacionada con la colisión definitiva de la Costa con la margen andina. La Costa emergió despues, hasta
la formación de nuevas cuencas de antearco en el Oligoceno terminal-Mioceno.
Las fases tectónicas del Paleoceno superior, Eoceno inferior y Eoceno superior basal son las más importantes
conocidas en la Costa ecuatoriana y traducen su acreción progresiva con la margen. La erosión tectónica parece
ser responsable de la creación repetida de cuencas de antearco, ya que cada fase de subsidencia sigue una fase
compresiva que, al favorecer la fricción en el plano de subducción, provocaría la erosión mecánica de la base de
la placa superior. Dicha evolución sedimentaria discontinua aclara las condiciones de enterramiento y madu-
ración de la materia orgánica, y la estructura geológica compleja que resultó debe ser tenida en cuenta para
futuros trabajos de exploración petrolera.

INTRODUCTION
Coastal Ecuador has been identified as an allochtho-
nous terrane of oceanic origin (Goossens and Rose, 1973;
Juteau et al., 1977; Lebrat et al., 1987), accreted to the
Andean continental margin during Late Cretaceous–
early Tertiary time (Feininger and Bristow, 1980;
Shepherd and Moberly, 1981; Lebrat et al., 1987). The
allochthonous nature of coastal Ecuador is supported by
a gravimetric survey (Feininger and Seguin, 1983) and by
paleomagnetic studies that show a 70° clockwise rotation
of this area has occurred since the middle Cretaceous
(Roperch et al., 1987). Since the Eocene, these regions of
accreted basement have remained in a fore-arc setting
(Figure 1) (Benitez, 1983; Mégard, 1987; Daly, 1989;
Marksteiner and Alemán, 1991).
In coastal Ecuador, two main zones have been recog-
nized. They are separated by the present-day Chongón-
Colonche fault which has been interpreted as a major
paleogeographic feature (Canfield, 1966; Benitez, 1983,
1992). North of the Chongón-Colonche fault, on the
Chongón-Colonche Cordillera and in the Manabí basin
(Figure 2), the stratigraphic succession is characterized
Figure 1—Location map of the southern coast of Ecuador.
by middle or upper Eocene beds unconformably
overlying the Cretaceous–lower Paleocene interval.
South of the Chongón-Colonche fault, the stratigraphic
succession of the Santa Elena Peninsula is characterized successive sedimentary basins. This study was part of a
by a thick upper Paleocene sequence and by the develop- scientific cooperative agreement between the Ecuadorian
ment of the deeply subsided Progreso basin of Neogene state oil company Petroecuador-Petroproducción and the
age (Figure 3). French Institute of Scientific Investigations for Develop-
The occurrence of oil in southern coastal Ecuador ment in Cooperation–ORSTOM.
motivated several geologic studies that have led to
numerous and often contradictory interpretations. In this
paper, we present a synthesis of the available strati- PREVIOUS WORK
graphic and sedimentologic data, as well as new field
work and paleontologic studies. These efforts have The discovery of small oil fields in the southwestern
refined the stratigraphic framework and have enabled us part of the Santa Elena Peninsula at the beginning of this
to revise the sedimentologic and paleogeographic inter- century led to detailed paleontologic and micropaleonto-
pretations and to modify earlier interpretations of the logic studies of the Upper Cretaceous and Tertiary
tectonic and sedimentary evolution of this area. This stratigraphy of southern coastal Ecuador (e.g., Sinclair
paper specifically addresses the Cenomanian–Eocene and Berkey, 1923; Olsson, 1931, 1942; Thalmann, 1946,
interval, during which the geologic history is marked by Cushman and Stainforth, 1951; Sigal, 1969) and estab-
changing paleotectonic settings and development of lished regional stratigraphic relationships (Sheppard,
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 617

companies, most of which is unpublished. In this study,


we draw upon both new field observations and an
extensive synthesis of biostratigraphic information.

LATE CRETACEOUS–LATE EOCENE


TECTONOSTRATIGRAPHIC EVOLUTION
Late Cretaceous–Early Late Paleocene
The basement of coastal Ecuador (Piñon Formation) is
made up of massive tholeiitic basaltic and basalt-
andesitic lavas generally considered to be a piece of
oceanic floor (Goossens and Rose, 1973; Juteau et al.,
1977; Feininger and Bristow, 1980; Mégard, 1987; Daly,
1989). However, early chemical studies showed that the
Piñon Formation has affinities with island arc volcanic
rocks (Goossens et al., 1977; Henderson, 1979). More
recently, Lebrat et al. (1987) distinguished the altered and
metamorphosed Piñon Formation of N-type MORB
composition, dated as late Aptian–Albian (110 ± 10 and
104 ± 15 Ma) (Goossens and Rose, 1973), and the San
Lorenzo Formation of island arc nature, which is late
Campanian–Paleocene in age and crops out only in the
Manabí area (see also Faucher and Savoyat, 1973;
Wallrabe-Adams, 1990). The Piñon Formation (Figure 3)
crops out in the Guayaquil area, in the Chongón-
Colonche Cordillera, and in the southwestern part of the
Santa Elena Peninsula and has been recognized in the
Manabí basin, thus suggesting that it constitutes the
basement of the entire southern coastal Ecuador
(Figures 3, 4).

Figure 2—Structural and morphologic setting of southern Cenomanian–Coniacian


coastal Ecuador, and location of the main localities cited in The Calentura Formation conformably overlies the
the text. Stars indicate oil fields. Piñon Formation (Figure 3) (Alvarado and Santos, 1983;
Benitez, 1990). It is a 200-m-thick succession of shales,
black laminated limestones, and thin-bedded graywacke
1937; Marchant, 1961; Sauer, 1965; Canfield, 1966; turbidites that were deposited in a pelagic, partially
Faucher et al., 1971; Faucher and Savoyat, 1973). How- anaerobic environment and that include a few thin-
ever, the subsurface studies carried out by different oil bedded volcanic breccias and hyaloclastites. The
companies in specific areas resulted in contradictory foraminifera indicate a late Cenomanian–Turonian age
local stratigraphic nomenclature and ages. (Thalmann, 1946; Sigal, 1969), which has been partially
With emergence of the plate tectonic theory, the appar- confirmed by the discovery of a Turonian ammonite (R.
ently confusing stratigraphy, the poor quality and scarcity Marocco, 1992, personnal communication). Nannofossils
of outcrops, and the tectonic complexity of the Santa indicate an early Coniacian age (Gamber et al., 1990). The
Elena Peninsula area led Azad (Anglo-Ecuadorian propri- Calentura Formation is known in the Guayaquil area and
etary report, 1968) and Colman (1970) to interpret the the eastern part of the Chongón-Colonche Cordillera, but
geology of the Peninsula as a giant olistostrome of late has not been recognized farther west. The Calentura
Eocene age and involving Upper Cretaceous–middle Formation is attributed to starved, deep-marine pelagic
Eocene rocks. This interpretation was shared by Bristow sedimentation deposited in Cenomanian–early
and Hoffstetter (1977) and Feininger and Bristow (1980). Coniacian time on the young Piñon oceanic crust
In the 1980s, studies carried out by Petroecuador and (Figure 4).
the Escuela Superior Politécnica del Litoral of Guayaquil
(Espol) restored the former stratigraphic framework and Santonian–Campanian
elaborated sedimentologic interpretations based mainly
on submarine fan models (e.g., Benitez, 1983, 1992; Egüez, Guayaquil Area The Cayo Formation conformably
1985; Nuñez del Arco et al., 1986; Santos et al., 1986a; overlies the Calentura Formation (Benitez, 1990; Mark-
Contreras, 1990). Meanwhile, a large amount of micropa- steiner and Alemán, 1991). It crops out on both sides of
leontologic and geologic work has been carried out by oil the Chongón-Colonche fault. The Cayo Formation is a
618 Jaillard et al.

Figure 3—Stratigraphic
framework of the sedimentary
units of southern coastal
Ecuador.

Figure 4—Stratigraphy and


environment of the Creta-
ceous–middle Paleocene
deposits of southern coastal
Ecuador in the Guayaquil area
(after Benitez, 1990; Mark-
steiner and Alemán, 1991) and
the Manabí area (after Faucher
et al., 1971).
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 619

2000-m-thick succession of fining-upward, coarse- ramis sp., Buryella aff. tetradica, Cenosphaera sp., Dicty-
grained volcaniclastic sandstones and conglomerates, omitra aff. andersoni, Lychnocanoma sp., Phormocyrtis striata
including a spectrum from high- to low-density exquisita, Protoxiphotractus sp., and Stylosphaera sp.
turbidites with shaly intercalations (Figure 4). Planktonic indicate a Paleocene age. These outcrops are altered,
foraminifera and dinocysts indicate a late Santonian– weakly metamorphosed (Sheppard, 1937), and intensely
Maastrichtian age (Thalmann, 1946; Bristow 1976; deformed, thus precluding any precise estimate of
Benitez, 1990; Gamber et al., 1990). Reworked foram- thickness or detailed sedimentologic analysis (Figure 5).
inifera indicate a shallow-marine provenance, and scarce However, the lithology is comparable to that of the
paleocurrent data suggest a west-directed transport Guayaquil Formation, suggesting a similar depositional
(Benitez, 1990). This coarse-grained sedimentation environment.
contrasts markedly with the underlying fine-grained The deformation of the Santa Elena Formation
deposits and indicates that an important tectonic and involves gently southward-dipping shear planes and
geodynamic change occurred by late Coniacian–early tight folds with ENE-WSW to WNW-ESE trending axes,
Santonian time. The Cayo Formation is attributed to associated with a penetrative axial plane cleavage
erosion of a volcanic terrain, which is thought to have dipping gently toward the south. The overall deforma-
been an island arc (Wallrabe-Adams, 1990; Marksteiner tion of the formation clearly increases toward the
and Alemán, 1991) (Figure 4). However, further southwest. The orientations of axial cleavages and shear
geochemical and mineralogic studies are necessary to planes, the warping of the folds, and the analysis of the
identify the provenance precisely. displacement criteria indicate a heterogeneous deforma-
tion associated with north- to NNW-directed thrust
Manabí Area Coarse-grained volcaniclastic conglom- vergence. Because it affects lower upper Paleocene beds
erates associated with basalt flows and dikes that were (Santa Elena Formation) and is covered by uppermost
known as the Cayo Formation (Faucher et al., 1971) are Paleocene coarse-grained deposits, this major tectonic
now referred to as the San Lorenzo Formation (Lebrat et event is of late Paleocene age (about 57 ± 2 Ma, according
al., 1987) (Figure 4). These beds rest on massive basalts to Haq et al., 1987).
ascribed to the Piñon Formation (Faucher et al., 1971).
They have yielded late Campanian radiolarians at Northeast of Chongón-Colonche Fault The little
Machalilla (Romero, 1990). In the Manta area, deformed Guayaquil Formation conformably and grada-
interbedded pillowed basalts as well as dikes and small tionally overlies the Cayo Formation (Figures 4, 5). It
plutons cross cutting the formation have yielded crops out only north of the Chongón-Colonche fault. It
Santonian–early Eocene radiometric ages (85–52.9 Ma), consists of about 400 m of dark, siliceous tuffs and shales,
with a maximum during the late Campanian–early with numerous cherts and subordinate thinly bedded
Maastrichtian (77–72 Ma) (Goossens and Rose, 1973; Hall turbidites, which contrast with the underlying coarse-
and Calle, 1982; Pichler and Aly, 1983; Lebrat et al., 1987; grained Cayo Formation. Thalmann (1946), Sigal (1969),
Wallrabe-Adams, 1990). These volcanic rocks have an and Faucher et al. (1971) identified planktonic
island arc composition and are much less altered than foraminifera of Maastrichtian age, with a probable
the rocks of the Piñon Formation (Lebrat et al., 1987). At extension into the Paleocene. This was confirmed by
Machalilla, the occurrence of numerous andesitic Benitez (1991), Gamber et al. (1990), and our work
boulders and clasts derived from the Cayo Formation because of nannofossils and radiolarians that indicate an
indicates that the deposition is coeval with synsedimen- early late Paleocene age (tympaniformis zone) for the top
tary tectonic deformation that caused subaerial erosion of of the formation near the town of Guayaquil.
the formation. Paleocurrents are locally directed toward The Guayaquil Formation is attributed to pelagic sedi-
the WSW (Romero, 1990). Chemical and geologic data mentation that was coeval with mild or distal volcanic
have shown that the San Lorenzo Formation can be activity (Figure 4). The lack of any significant quartz-rich
attributed to erosion of an active island arc (Lebrat et al., detritus suggests either that southern coastal Ecuador
1987; Marksteiner and Alemán, 1991). was located far from a continental source or that the area
was sheltered from any significant continental detrital
Maastrichtian–Early Late Paleocene supply. The increasing amount of calcareous nodules or
South of Chongón-Colonche Fault The Santa Elena beds toward the top of the unit suggests a slight shal-
Formation crops out only in the Santa Elena Peninsula lowing-upward trend (Benitez, 1991) and possibly a
(Figures 3, 5). It has long been considered a stratigraphic deepening of the carbonate compensation depth (CCD).
equivalent of the Guayaquil Formation (Sinclair and The late Paleocene tectonic phase that followed the depo-
Berkey, 1923; Thalmann, 1946; Canfield, 1966). We have sition of the Santa Elena Formation ended this phase of
recently confirmed this interpretation. Some outcrops on sedimentation.
the peninsula have yielded the radiolarians Amphy-
pyndax tylotus, Archaeodictyomitra lamellicostata, Diacantho- Northwest of the Chongón-Colonche Fault In the
capsa granti, and Stylospongia sp., among others, and the Manabí area, coarse-grained graywackes intercalated
calcareous nannofossils Arkhangelskiella cf. scapha, Coccol- with basaltic flows and ash beds yield Maastrichtian–
ithus paenepelagicus, Micula decussata, Quadrum gartneri, Paleocene(?) microfaunas (Sigal, 1969; Faucher et al.,
and Watznaueria barnesae, which indicate a latest Creta- 1971), which are consistent with the Maastrichtian–
ceous age. In other samples, the radiolarians Bathropy- Paleocene radiometric ages obtained from the top of the
620 Jaillard et al.

Figure 5—The Upper Creta-


ceous–upper Eocene strati-
graphy of the Santa Elena
Peninsula and Chongón-
Colonche Cordillera.

San Lorenzo Formation (Hall and Calle, 1982; Lebrat et and shales (Bristow and Hoffstetter, 1977). These
al., 1987; Wallrabe-Adams, 1990). Therefore, magmatic sediments were deposited on submarine fans largely by
activity related to an island arc went on in this area while high-density turbidites, with a minor amount of low-
the Guayaquil and Santa Elena formations were being density flows (Moreno, 1983; Benitez, 1983). Various
deposited (Figure 4). This observation, together with the formations have been recognized (Marchant, 1961; Small,
lack of Cenomanian–Coniacian deposits in this area, 1962; Canfield, 1966). However, their stratigraphic
suggests that the volcaniclastic sedimentation is diachro- succession is not established and they cannot be used for
nous and the volcanic activity migrated from the mapping purposes, thus detracting from their usefulness
Guayaquil area in the early Late Cretaceous and toward (Benitez, 1992).
the Manabí area in the latest Cretaceous–Paleocene. On the basis of benthonic foraminifera, the Azúcar
Group has long been considered early Paleocene in age
(Thalmann, 1946; Small, 1962; Benitez, 1992; Marksteiner
Late Paleocene–Early Eocene and Alemán, 1991), although the mollusk fauna suggests
a younger age (Olsson, 1942; Canfield, 1966; Sigal, 1969;
Late Paleocene Faucher et al., 1971; Daly, 1989). In contrast, the plank-
The conspicuous Azúcar Group is known only south tonic foraminifera (e.g., Globigerina cf. velascoensis, G. trilo-
of the Chongón-Colonche fault (Figures 3, 5). Although culinoides, G. aff. daubjergensis, Globorotalia angulata and G.
the lower contact has not been observed, it is most mackannai, Small, 1962; Moreno, 1983; Litton Resources
probably unconformable on the Santa Elena Formation, proprietary report, 1986; Gamber et al., 1990) are of
as suggested by the analysis of seismic lines (Marksteiner middle–late Thanetian age (pseudomenardii and velas-
and Alemán, 1991). The Azúcar Group consists of at least coensis zones), and indicate that most of the benthonic
1500 m of conglomerates, pebbly sandstones, sandstones, foraminifera are reworked (Figure 5).
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 621

In the southern part of the Santa Elena Peninsula,


conglomeratic clasts are mainly derived from the Santa
Elena and Guayaquil formations, continental basement,
and volcanic rocks (Marksteiner and Alemán, 1991). This
indicates that the Cretaceous strata were deformed and
subjected to substantial erosion and that the Santa Elena
Peninsula was in contact with the Andean continental
margin. Paleocurrents indicate a north-northeastward
transport direction (Moreno, 1983) (Figure 6). In contrast,
in the northern Santa Elena Peninsula, the amount of
volcanic clasts is much greater (Marksteiner and Alemán,
1991) and preliminary results suggest southwest-directed
paleocurrents. The drastic change of provenance with
respect to older deposits is clearly related to the major
phase of late Paleocene deformation, and the Azúcar
Group postdates this event.
The Azúcar Group is also well structured, including
faulting and ENE-WSW-trending tight folds with vertical
axial planes. This tectonic phase is assigned to the early
Eocene because similar deformation is not present in the
overlying Lutetian sequence. As a consequence, the
observed deformation of the Santa Elena Formation
Figure 6—Paleocurrents measured in the Azúcar Group
apparently resulted from superimposition of an early
(latest Paleocene) in the Playas area (Santa Elena
phase characterized by ESE-WNW-trending, warped Peninsula). (After Moreno, 1983.)
tight folds of late Paleocene age, which affected only the
Santa Elena Formation, with a later ENE-WSW-trending
deformation which affected the Santa Elena Formation as
represent the lenticular “basal conglomerate” of the
well as the Azúcar Group.
overlying mainly Lutetian sedimentary cycle.
The Early Eocene Problem
North of Chongón-Colonche Fault West of
The early Eocene was marked by a widespread sedi- Guayaquil on the southern side of the Chongón-
mentary hiatus. Although no formations of this age are Colonche Cordillera, planktonic foraminifera and
known in southern Ecuador, some studies report early calcareous nannofossils of latest Paleocene–early Eocene
Eocene fossils from poorly known beds (Figure 3 and 5). age have been recognized in limestone samples (Unocal
proprietary report, 1987), which probably correspond to
South of Chongón-Colonche Fault In the Santa Elena the base of the San Eduardo Formation. In both areas, the
Peninsula, the “Passage beds” and equivalent units were lower Eocene sedimentation is either condensed or
only recognized in well cutting samples; there are no thinned by erosion. These characteristics suggest that the
direct data of a sedimentologic nature. These beds, the early Eocene hiatus resulted from widespread, possibly
thickness of which varies up to 350 m, have a lenticular diachronous emergence that occurred between earliest
shape and consist of two distinct horizons. The first Ypresian and earliest Lutetian times.
comprises shales and micaceous sandstones bearing
planktonic foraminifera of early Eocene age (Globigerina
aff. stonei, Globorotalia aff. acuta, G. aff. aequa, G. crassata) Late Early Eocene–Early Middle Eocene
which are associated with reworked benthonic (Late Ypresian–Lutetian)
foraminifera (Thalmann, 1946; Small, 1962; Bristow and
Hoffstetter, 1977). The second layer includes locally Late Ypresian–early Lutetian time was characterized
conglomeratic sandstones and shales with plant remains by a widespread transgression associated with tectonic
and is characterized by species of Discocyclina (Barker, subsidence, which allowed deposition of a thick, shal-
1932; Anglo-Ecuadorian Oilfields proprietary report, lowing-upward marine sequence. North of the
1956), suggesting a correlation with the lowest Lutetian Chongón-Colonche fault, the transgression is markedly
beds of northwestern Peru which contain a similar diachronous; it overlies deeply eroded rocks and is asso-
foraminiferal fauna (González, 1976). ciated with conspicuous synsedimentary tectonism. This
The “lower Passage beds,” containing mainly plank- diachronism, together with the variable facies, has
tonic foraminifera, may represent the end of the resulted in numerous poorly defined stratigraphic units.
preserved remnants of the uppermost Paleocene marine
Late Early Eocene–Middle Lutetian
sedimentary cycle (Azúcar Group). In the “upper
Transgression
Passage beds,” which are apparently absent north of the
Chongón-Colonche fault, the presence of plant fragments South of Chongón-Colonche Fault In the Ancón area,
and benthonic foraminifera indicates a shallow marine the Clay Pebble beds overlie the upper Passage beds. The
environment. The upper Passage beds may therefore Clay Pebble beds consist of up to 700 m of disrupted
622 Jaillard et al.

(Figures 5, 7). West of Guayaquil, the base of the


formation consists of a few meters of bedded cherty
marls and shales (Figure 8). We have identified radio-
larians (Lamptonium cf. fabaeforme, Orbula discipulus, Phor-
mocyrtis striata exquisita), calcareous nannofossils (Fasci-
culithus tympaniformis, Heliolithus kleinpelli, H. cf. riedelli,
Tribrachiathus orthostylus), and planktonic foraminifera
(Globigerina aff. primitiva, G. aff. collactea, Globorotalia
aequa, G. broedermanni, G. esnaensis, G. pseudotopilensis, G.
wilcoxensis), which range from late Paleocene to late early
Eocene in age. These are associated with reworked Creta-
ceous benthonic foraminifera. We interpret the base of
the San Eduardo Formation as middle–late Ypresian in
age (aragonensis zone). The overlying calciturbidites
contain numerous algae, oncolites, and benthonic
foraminifera (Discocyclina and Asterocyclina). Together
with associated calcareous nannofossils and planktonic
foraminifera, they indicate an earliest Lutetian age
(Bristow and Hoffstetter, 1977; Gamber et al., 1990). At
the top of the formation, we found radiolarians
(Podocyrtis aff. diamesa, Thyrsocyrtis hirsuta) of late
Ypresian–early Lutetian age (P 9–10 zones). In the
western part of the Manabí basin, the San Eduardo
Formation unconformably overlies Upper Cretaceous
rocks, whereas in the eastern part of the Chongón-
Colonche Cordillera, it conformably overlies the
Paleocene Guayaquil cherts, thus suggesting a westward
increase of pre-Lutetian erosion (Figure 8).
In the western part of the Chongón-Colonche
Cordillera northeast of Colonche (Figure 2), the brec-
ciated Guayaquil Formation is overlain by thin lower
Lutetian marls. These in turn are overlain by a few
Figure 7—Stratigraphic correlations of the Lutetian trans- meters of fining-upward calcareous conglomerates that
gressive deposits in the Chongón-Colonche Cordillera. reworked oncolitic limestones, and some Maas-
trichtian–Paleocene cherts (Javita limestones) (Sigal, 1969;
Benitez, 1992) (Figures 7, 8).
shales, including clasts and contorted beds of pebbly In most of the Manabí basin, the Lutetian transgres-
sandstones, sandstones, shales, cherts, and limestones sion is reflected in radiolarian-bearing, fine-grained
(Figure 5). They represent large-scale slumps (Brown and shales, tuffaceous cherts, and siliceous limestones of
Baldry, 1925; Marchant and Black, 1960; Marksteiner and pelagic origin that are dated as middle Lutetian (Cerro
Alemán, 1991) that express instability of the substratum. Formation, base of the San Mateo and Punta Blanca
The Clay Pebble beds do not constitute a formation, but formations) (Sigal, 1969; Bristow and Hoffstetter, 1977;
rather a diachronous facies that occurs at the base and Romero, 1990). These beds either conformably overlie the
within the lower part of the Lutetian sequence (Anglo- San Eduardo Formation or unconformably rest on the
Ecuadorian Oilfields proprietary report, 1956). In the Senonian San Lorenzo Formation (Figure 8). They
Ancón oil field, its age ranges from latest Ypresian to commonly contain breccias, reworked Cretaceous micro-
early Lutetian (Bristow and Hoffstetter, 1977; Jiménez fauna, and olistoliths of Cretaceous rocks (Schulman et
and Mostajo, 1990). In other regions of the peninsula, the al., 1965), which indicate that the early–middle Lutetian
Lutetian beds unconformably overlie Cretaceous or transgression was associated with tectonism and erosion.
lower Tertiary rocks and contain at the base a 0- to 30-m-
thick, coarse-grained basal conglomerate of early–middle Lutetian Sequence
Lutetian age (Small, 1962; Bristow and Hoffstetter, 1977; After the diachronous and tectonically driven trans-
Rosario conglomerate of Benitez, 1992). gression of late Ypresian–middle Lutetian age, the rest of
the Lutetian corresponds to a shallowing-upward
North of Chongón-Colonche Fault North of the sequence of marine shelf deposits (Figure 5).
Chongón-Colonche fault, the early Lutetian transgres-
sion is generally expressed by the San Eduardo South of Chongón-Colonche Fault In the Santa Elena
Formation which consists of 30–120 m of well-stratified Peninsula, the Lutetian sequence is 1000–1500 m thick
calciturbidites deposited within autochthonous, and comprises the Clay Pebble beds and the Socorro and
hemipelagic marls and micrites (Santos et al., 1986b) Seca formations (Anglo-Ecuadorian Oilfields proprietary
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 623

Figure 8—Chronostrati-
graphic sketch of the middle
Ypresian–Lutetian trans-
gression along the
Chongón-Colonche
Cordillera.

report, 1956) (Figure 5). The Socorro Formation consists secondary gypsum veinlets, and heavy mineral laminae
of laminated shales, siltstones, and fine-grained sand- and exhibits characteristics typical of clastic shore zone
stones of an outer shelf environment, intercalated with sequences (shoreface to foreshore). Although no
some thick-bedded turbiditic sandstones. Slumped beds complete section has been studied, the depositional envi-
(including Clay Pebble facies) and turbidites are ronment and evolution are thought to be comparable to
common near the base and decrease upward, suggesting that of the Socorro and Seca formations of the Santa
a decrease in tectonic activity. The Socorro Formation Elena Peninsula. In the Manabí basin, undated layers of
grades upward into the Seca Formation, a sequence of the San Mateo Formation locally rest on volcanic rocks
laminated shales, siltstones, and marls that reflect ascribed to the Lower Cretaceous Piñon Formation
climatic or seasonal influences, as well as thin-bedded (Figure 8).
sandstones attributed to storm processes and subordi-
nate turbidites (Figure 9). In the Seca Formation, the Middle–Late Eocene (Bartonian–Early
upward increase of bioturbation, calcareous content, and Priabonian)
neritic fauna indicate a shallow shelf environment. The
Socorro and Seca formations contain calcareous nanno- In southern coastal Ecuador, the middle Eocene
fossils, planktonic foraminifera, radiolarians, mollusks, sequence ends with continental to shallow marine
and reworked benthonic foraminifera in the turbiditic coarse-grained graywackes and lithic sandstones that
beds, which together indicate an early–late Lutetian age abruptly overlie the Lutetian marine sequence. These
(Bristow and Hoffstetter, 1977; Jiménez and Mostajo, deposits are called the Punta Ancón Formation along the
1990; Gamber et al., 1990). Sedimentary measurements present-day coast of the Santa Elena Peninsula and the
indicate NNW- to WNW-directed paleocurrents and a San Mateo Formation (upper part) in the Manabí basin
northwest-dipping paleoslope (Figure 9). (Figures 5, 8). On the inner part of the Santa Elena
Peninsula and in the Chongón-Colonche Cordillera, the
North of Chongón-Colonche Fault In the southern so-called Zapotal Formation apparently comprises two
part of the Chongón-Colonche Cordillera, the 350-m- stratigraphic units. One consists of coarse-grained,
thick Lutetian sequence is known as the Las Masas poorly dated, continental to coastal deposits with
Formation. In the Manabí basin, contemporaneous beds molluscan fauna (Hannatoma fauna) and rare marine
are 500–1500 m thick and correspond to the lower part of microfauna that broadly correlate with the lower
the San Mateo Formation (Figures 5, 8). The foraminiferal Priabonian beds of northern Peru (Verdún Formation)
and radiolarian content indicate a Lutetian–early late (Olsson, 1931; Paredes, 1958; González, 1976; Bristow
Eocene age for the entire San Mateo Formation and Hoffstetter, 1977). The other unit consists of fine-
(Cushman and Stainforth, 1951; Sigal, 1969; Bristow and grained, clastic marine deposits dated as late
Hoffstetter, 1977; Navarrete, 1986; Contreras, 1990). Oligocene–early Miocene by planktonic foraminifera,
However, the Bartonian and late Eocene faunas were suggesting that it belongs to the overlying mainly
probably found in the coarse-grained upper part of the Neogene sedimentary cycle (Bristow, 1975; Bristow and
formation. The Las Masas and lower San Mateo forma- Hoffstetter, 1977). We agree with Olsson (1931), Canfield
tions are made up of partially calcareous shales, silt- (1966), and Sigal (1969) that the coarse-grained lower
stones, sandstones, and graywackes. South of Puerto part of the Zapotal Formation is partially equivalent to
Cayo, thin-bedded turbidites, tempestites, and rippled the Punta Ancón Formation of Bartonian–early
beds indicate a shelf environment that was shallower Priabonian age. This implies that a major sedimentary
than that of the Cerro Formation. In San Mateo west of hiatus of Oligocene age separates the lower and upper
Manta, the 700-m-thick Lutetian–Bartonian San Mateo parts of the Zapotal Formation.
Formation (Contreras, 1990) includes plant fragments, We are able to confirm the Bartonian age of the Punta
624 Jaillard et al.

Figure 10—Stratigraphic section of the Punta Ancón


Formation (Bartonian) in the Punta Ancón beach area
(Santa Elena Peninsula).

and coarsening-upward clastic sequences of shore zone


origin (Figure 10). Locally, massive conglomeratic facies
(west of Ancón, east of Playas) are interpreted as large
fluvial or distributary channels that fed into the clastic
coastal system. Measurements indicate southwest- to
northwest-directed paleocurrents perpendicular to the
present-day coast (see Figure 12). The amount of
conglomerates markedly increases northward, and in the
Colonche area, the Punta Ancón Formation grades
eastward into the lower Zapotal Formation.
The lower Zapotal Formation is a 300–600 m thick
succession of lithic sandstones and coarse-grained
conglomerates deposited in an alluvial environment. An
intermediate layer contains plant-bearing shales and
subordinate sandstones (Small, 1962; Canfield, 1966).
These argillaceous beds are believed to correlate with the
Figure 9—Stratigraphic section and paleocurrent data for upper Eocene Jusa Formation which is exposed 15 km
the Socorro and Seca formations (Lutetian) in the Punta ESE of Colonche (Cushman and Stainforth, 1951; Bristow
Ancón beach area (Santa Elena Peninsula). and Hoffstetter, 1977). In the Chongón-Colonche
Cordillera, the lower Zapotal Formation overlies either
Upper Cretaceous–Paleocene or middle Eocene deposits
Ancón Formation (Bristow and Hoffstetter, 1977; Jiménez and consists of coarser grained alluvial fan deposits.
and Mostajo, 1990) on the basis of rich radiolarian associ- These grade southward into finer grained deposits of
ations. In the southwestern part of the Santa Elena alluvial plain or coastal environment in the Santa Elena
Peninsula, the formation consists of reddish shales and Peninsula. At the southeastern end of the Peninsula
siltstones, lithic sandstones, and subordinate conglomer- (Posorja, Figure 2), the lower Zapotal Formation consists
ates, all of which are organized in typically thickening- of coastal sandstones similar to those of the Punta Ancón
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 625

Figure 12—Paleocurrents and paleogeographic interpreta-


tion of the Bartonian–early Priabonian deposits of southern
Figure 11—Stratigraphic column of the San Mateo coastal Ecuador.
Formation (middle–upper Eocene) in the Julcuy area.
(Simplified after Egüez, 1985.)

Formation and contains the Hannatoma molluscan fauna tion (lower Zapotal and upper San Mateo formations)
of early late Eocene age (Olsson, 1931; González, 1976). and reworking of Cretaceous and Paleogene rocks. These
North of Manglaralto, the Punta Ancón Formation coarse-grained deposits grade westward (Manabí
grades northward into conglomerate-prone deposits that present-day coast) into fan delta deposits (upper San
correspond to the upper part of San Mateo Formation of Mateo Formation) and south- or southwestward into
Bartonian–early Priabonian age (Cushman and Stain- alluvial plain systems (lower Zapotal Formation) and
forth, 1951; Sigal, 1969; Navarrete, 1986; Contreras, 1990). coastal deposits (Punta Ancón Formation) (Figure 12).
Along the present-day coast, it consists of a few hundred This paleogeographic setting, together with the volcanic-
meters of coarse-grained conglomeratic lenses and beds rich nature of the deposits (Figure 13) and the locally
of alluvial origin intercalated within the shoreline important pre-Bartonian unconformity, clearly indicate
sandstone sequence, indicating a fan delta depositional that the Chongón-Colonche Cordillera was drastically
setting. Farther east and southeast in the Julcuy area, the rejuvenated near the Lutetian–Bartonian boundary and
San Mateo Formation consists of a 600-m-thick sequence submitted to intense erosion. Late Priabonian–late
of coarse-grained conglomerates similar to the lower Oligocene time is characterized by a widespread sedi-
Zapotal Formation, with imbricated polymictic clasts, mentary hiatus (Canfield, 1966; Sigal, 1969; Bristow and
debris flows, and olistoliths apparently deposited in an Hoffstetter, 1977; Benitez, 1992).
alluvial fan environment (Figure 11). There, the San Consequently, the Bartonian–early Priabonian time
Mateo Formation generally rests directly on Cretaceous span (about 42–38 Ma, after Haq et al., 1987) is inter-
rocks (Cayo, San Lorenzo, or Piñon formations), indi- preted as a period of pronounced tectonic activity that
cating strong pre-Bartonian erosion and a conspicuous culminated in emergence of the entire area during the
basal unconformity. In the entire Manabí area, paleocur- late Priabonian. In the Santa Elena Peninsula, the late
rents indicate a NNW- to northwest-oriented transport Eocene deformation resulted in open folds trending
(Egüez, 1985; Santos et al., 1986a; Contreras, 1990) north-south to northeast-southwest associated with east-
(Figure 12). southeast gently dipping reverse faults, which indicate a
In summary, the Bartonian–Priabonian paleogeo- grossly ESE-WNW compression associated with WNW-
graphy comprises a central area (Chongón-Colonche ward thrust movements. Such deformation has not been
Cordillera, Manabí hills) marked by alluvial fan deposi- observed in the Neogene deposits.
626 Jaillard et al.

clastites at the base of the sequence near Guayaquil, and


the apparent lack of Cenomanian–Coniacian deposits
(Calentura Formation) in the Manabí region are consis-
tent with this interpretation.

Accretion and Early Fore-Arc Basin Stage:


Late Paleocene–Early Eocene
The late Paleocene–early Eocene was marked by the
occurrence of major tectonism that caused drastic
changes in the paleogeography (Figure 5, 14). The nature
of the coarse-grained detrital deposits indicates that
sialic, sedimentary, and volcanic provenance areas were
intensely deformed and deeply eroded. The differences
in the tectonic and sedimentary evolutions of the Santa
Elena Peninsula and the Chongón-Colonche-Manabí
area indicate that they represent independent structural
units during the Paleocene–early Eocene deformation
phases (Figure 14).
In the Santa Elena Peninsula, intense deformation of
the middle Paleocene Santa Elena Formation is followed
Figure 13—Nature of the detritus during the early middle by substantial tectonic subsidence that accommodated
Eocene (Socorro and Seca formations) and the late middle deposition of thick uppermost Paleocene coarse-grained,
and late Eocene (Punta Ancón and lower Zapotal forma- quartz-rich turbidites (Azúcar Group) (Figure 14). North
tions). (After Marksteiner and Alemán, 1991.) of the Chongón-Colonche fault, this tectonic phase is
believed to have induced the sedimentary hiatus of latest
Paleocene–early Eocene age. Although the early Eocene
period is still poorly understood, it is apparent that the
TECTONISM AND BASIN Santa Elena Peninsula became emergent at this time.
DEVELOPMENT These events are interpreted as the result of collision of
the Cayo remnant arc with the continental margin, which
The Late Cretaceous–late Eocene evolution of coastal provoked the blocking of the subduction and probably
Ecuador reflects three phases that were separated by the thrusting of the Chongón-Colonche Cordillera and
major tectonosedimentary events (Figure 14). Manabí areas (Figure 14). The shallowing-upward
sequence of latest Paleocene–early Eocene age represents
Marginal Basin Stage: Late the infilling of the first real fore-arc or slope basin in
Cretaceous–Middle Paleocene southern coastal Ecuador history.
In the Talara basin, which formed above the Amotape
Between the Late Cretaceous and middle Paleocene, massif of northwestern Peru, the Paleocene–Eocene
volcaniclastic pelagic sediments accumulated on the boundary coincides with a thick unconformity-bound
Early–middle Cretaceous oceanic floor of southern succession of continental derived sandstones and
coastal Ecuador. The lack of significant quartz-rich conglomerates (Mogollón Formation) (González, 1976;
detritus suggests that they were not deposited from a Macharé et al., 1986; Séranne, 1987). This unconformity
sialic landmass. The presence of a thick coarse-grained clearly expresses a second important tectonic event of
sequence of Santonian–Campanian age (Cayo earliest Eocene age that we correlate with the wide-
Formation) suggests that the basin was bordered by an spread sedimentary hiatus observed in southern coastal
island arc, active at least since the Coniacian. The fining- Ecuador. These accretionary events, of late Paleocene
upward trend of this sequence indicates that the activity and earliest Eocene age, mark the end of the island arc
of the Cayo arc decreased with time. In contrast, volcanic marginal basin evolution in southern coastal Ecuador.
and tectonic activity since the late Campanian in the
Manabí area (San Lorenzo Formation) is interpreted as Fore-Arc Basin and Definitive Collision:
resulting from the formation of a new island arc. Middle–Late Eocene
Therefore, we interpret the Late Cretaceous–middle
Paleocene sequence as the infilling of a marginal basin The middle–upper Eocene sedimentary strata of
(Karig and Moore, 1975) opened between the early Late marine shelf and terrestrial origin are characterized by a
Cretaceous Cayo arc and the San Lorenzo island arc marked shallowing- and coarsening-upward trend
active in latest Cretaceous–early Paleocene time (Figure (Figures 5). This period ended with a major sedimentary-
14). According to the models of Karig and Moore (1975) stratigraphic hiatus, which encompassed most of the
and Carey and Sigurdsson (1982), the significant alter- Oligocene.
ation and metamorphism of the Early Cretaceous The early Lutetian transgression is attributed to exten-
volcanic rocks (Piñon Formation), the presence of hyalo- sional processes. This interpretation is supported by
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 627

Figure 14—Schematic cross sections showing the


evolution of southern coastal Ecuador from Late Creta-
ceous to Miocene time. (Rotations are not taken into
account.) (a) In the late Cretaceous, a marginal basin
opened between the early Late Cretaceous Cayo arc and
the latest Cretaceous–Paleocene San Lorenzo island arc.
(b) In the late Paleocene–earliest Eocene, the Cayo
remnant arc collided with the Andean continental margin
and caused intense deformation of the Santa Elena
Peninsula, emergence of the Chongón-Cordillera, and in-
filling of the Santa Elena basin by coarse-grained
turbidites. (c) Middle Eocene time was characterized by the
extensional subsidence of fore-arc basins, which were
deformed and probably inverted during the late Eocene
compressional phase. (d) Following a general emergence
phase of latest Eocene–Oligocene age, new extensional
fore-arc basins were formed as the Gulf of Guayaquil
opened.

by the Chongón-Colonche Cordillera swell. These fore-


arc basins were filled by the middle–upper Lutetian shal-
lowing-upward sequence.
The paleogeographic change expressed by the basal
unconformity and the volcanic-rich and coarse-grained
nature of the Bartonian–Priabonian deposits are inter-
preted as the result of definitive collision of southern
coastal Ecuador with the Andean continental margin. The
WNW-ESE compression determined for the late Eocene
tectonic phase is consistent with this interpretation.
Coarse-grained Priabonian deposits comparable to
those of southern coastal Ecuador occur in the
surrounding areas. In northwestern Peru, unconformable
lower upper Eocene conglomerates (Verdún Formation)
overlap Tertiary sedimentary strata of the Talara basin
(Paredes, 1958; González, 1976; Séranne, 1987) and the
Paleozoic–Cretaceous rocks of the broader area (Caldas
et al., 1980; Reyes and Caldas, 1987). In the Eastern
Cordillera of Ecuador, which constitutes the eastern edge
of the microplate of coastal Ecuador (Santos and
Ramírez, 1986), the middle Eocene succession ends in
coarse-grained deposits (Apagua conglomerates) inter-
preted as the result of collision and underthrusting of
this unit under the continental margin (Egüez and
Bourgois, 1986; Bourgois et al., 1990). This interpretation
matches our own observations.

New Fore-Arc Basin:


Late Oligocene–Miocene
The late Oligocene–earliest Miocene to Pliocene
interval is marked by the development of several basins
that were filled by fine-grained shallow marine sand-
persistent reworking and slumping, pronounced stones and shales (Figure 14). The creation of these
diachronism of the transgression, local pre-Lutetian Neogene troughs is thought to have been triggered by
erosion, and upward disappearance of the synsedimen- the opening of the Gulf of Guayaquil along the
tary deformation. This extensional driven transgression Guayaquil-Dolores megashear (Shepherd and Moberly,
following the early Eocene compression marks a 1981). In the Santa Elena Peninsula, the Neogene
renewed stage of fore-arc basin subsidence in southern Progreso basin is confined by a strand of the Chongón-
coastal Ecuador (Figure 14). The differences in the Colonche fault (Carrizal fault) and by the La Cruz fault
Lutetian evolution and lithologies observed on either (Figure 2), probably inherited from the late Paleocene–
side of the Chongón-Colonche fault suggest that the early Eocene collision (Figure 14). Their evolution is
Santa Elena basin was separated from the Manabí basin broadly contemporaneous with that of the Andean inter-
628 Jaillard et al.

montane sedimentary basins controlled by a transten-


sional-transpressional tectonic regime (Lavenu et al.,
1992; Marocco et al., 1995). Most of coastal Ecuador
became emergent during the Quaternary.

CONCLUSIONS
Between the Late Cretaceous and late Eocene, the
oceanic-floored allochthonous terranes of southern
coastal Ecuador underwent a complex geologic evolution
that included island arc related and marginal basin sedi-
mentation, collisions associated with prominent shear
deformation, basin subsidence, and several phases of
uplift. The Cenomanian–middle Paleocene phase was
characterized by pelagic and volcaniclastic sedimenta-
tion in a marginal basin that was remote from silicic
detrital influx. Late Paleocene–early Eocene time was
marked by intense tectonic deformation, voluminous
silicic detrital influx, and prominent sedimentary breaks
and ended with widespread emergence of coastal
Ecuador. These phenomena are attributed to collision of
southern coastal Ecuador with the Andean continental
margin, which caused the creation of an early, short-
lived fore-arc or slope basin. Middle Eocene time began
with a tectonically induced diachronous transgression.
This was followed by a middle Eocene shallowing- Figure 15—Estimated sedimentation rates in southern
upward sequence that represents the infilling of a new, coastal Ecuador during Late Cretaceous–late Eocene time.
short-lived fore-arc basin. Late Eocene is marked by
locally unconformable, coarser grained, shallow-marine
and continental deposits. The latest Eocene emergence of
southern coastal Ecuador is attributed to a major “can presumably form structural basins within which
compressive event preceding the development of a new substantial thickness of shelf or slope deposits can accu-
stage of fore-arc basin subsidence in the latest mulate.” Southern coastal Ecuador exhibits such a
Oligocene–Miocene. complex structural and sedimentary evolution, which
The late Paleocene and early Eocene tectonic phases can be explained by a multiple accretionary history
recorded in the Santa Elena Peninsula are undoubtedly (Figures 14). The hypothesis, according to which
the most intense deformation episodes undergone by mélange or a giant olistostrome formed during Paleo-
coastal Ecuador. The first was responsible for numerous gene time (Colman, 1970), has not been confirmed by our
gently dipping shear planes, subisoclinal folds, and study.
pervasive cleavage. The second apparently formed tight One striking feature of this evolution is the close asso-
vertical folds and faults. In contrast, the present-day ciation between phases of uplift and the succeeding
structure of the middle Eocene beds displays only subsidence, which created successive short-lived fore-arc
reverse faults and gentle folding with dips usually less basins separated by periods of erosion (Figure 15). This
than 30°, which are probably due mainly to the late suggests the existence of a genetic link between compres-
Eocene tectonic phase. From the late Eocene onward, the sive tectonic phases responsible for uplift and the subse-
Andean deformation shifted eastward in the present-day quent subsidence associated with extensional tectonism.
Andes toward the sub-Andean zone. However, the Such a relationship has been noted in the present-day
effects of these early tectonic phases have been largely margins of Japan and Peru and is believed to be due to
ignored or at least underestimated in the eastern regions. tectonic erosion of the edge of the continental margin
Except for Feininger and Bristow (1980), Roperch et al. along the plane of subduction (Von Huene et al., 1985;
(1987), and Lebrat et al. (1987), southern coastal Ecuador Suess et al., 1988; Von Huene and Lallemand, 1990). It is
has generally been analyzed and interpreted from an inferred that during the compressive phases responsible
autochthonist point of view. The recognition of for emergence, coupling between the oceanic and conti-
numerous tectonic events during the Late Cretaceous– nental lithospheres increased substantially, thus
late Eocene clearly indicates, however, that this was a provoking tectonic erosion of the lower surface at the
very mobile zone that underwent various types of trans- edge of the continental margin. This loss of mass along
lation and rotation (Roperch et al., 1987). Scholl et al. the continental margin was sufficient to initiate tectonic
(1980, p. 568) noted that “lateral tectonic accretion can subsidence of the fore-arc zone and the creation of
create a geographically wide ocean margin of exceptional subsiding fore-arc basins as the compressive stresses
structural, lithologic and stratigraphic complexity,” and decreased (Von Huene and Scholl, 1991).
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 629

Southern coastal Ecuador is an example of small oil experience of Andean and active margin geology during joint
fields located in an accretionary fore-arc setting. The field work. The final content of this manuscript was consider-
origin and evolution of the organic matter was controlled ably improved by a thorough review by D. Scholl and stimu-
by the geologic history. The organic-rich, fine-grained lating discussions with G. Mascle.
deposits of Cenomanian–Coniacian and Maas-
trichtian–middle Paleocene age (Calentura and
Guayaquil formations) form good potential source rocks REFERENCES CITED
for oil generation (Alvarado and Santos, 1983; Petro-
ecuador proprietary report) and have probably sourced Alvarado, G., and M. Santos, 1983, El miembro Calentura y la
the oil fields of southern coastal Ecuador. These sedimen- Formación Cayo: Actas del III Congreso Ecuatoriano Inge-
tary strata were deposited in an anaerobic marginal basin nieria Geología Minería y Petróleo, Guayaquil, v. I.A., 18 p.
isolated from silicic detrital influx, characterized by weak Barker, R. W., 1932, Larger foraminifera from the Eocene of
subsidence and low sedimentation rates. Santa Elena Peninsula, Ecuador: Geological Magazine, v.
All the producing oil fields in southern coastal 69, p. 302–310.
Ecuador occur on the Santa Elena Peninsula (Figure 2) Benitez, S., 1983, Contribución al estudio de las cuencas sedi-
where there was intense deformation and substantial mentarias del Suroeste ecuatoriano: Actas del III Congreso
latest Paleocene sedimentation. The late Paleocene Ecuatoriano de Ingenieria, Geología, Minería y Petróleo,
collision of part of southern coastal Ecuador against the Guayaquil, v. I.A., 41 p.
Benitez, S., 1990, Estratigrafía de las formaciones Cayo y
Andean continental margin deformed the stratigraphic Guayaquil en la Cordillera Chongón-Colonche: hacia una
succession and allowed deposition of thick, quartz-rich redefinición. Parte I: Geociencia, Guayaquil, v. 3, p. 7–11.
arenites in the Santa Elena Peninsula (Figure 14). These Benitez, S., 1991, Estratigrafía de las formaciones Cayo y
deposits may have favored burial and maturation of the Guayaquil en la Cordillera Chongón-Colonche: hacia una
organic matter, since the occurrence of oil fields in redefinición. Parte III: Geociencia, Guayaquil, v. 5,
southern coastal Ecuador coincides with the paleogeo- p. 11–14.
graphic extent of the upper Paleocene turbidites. Benitez, S., 1992, Estratigrafía del Paleógeno en el Ecuador:
Moreover, in the Ancón oil fields, the uppermost Simposium Nacional: Investigación y Desarrollo
Paleocene beds, which are called the Atlanta sandstone, tecnológico en el area de hidrocarburos, Conuep-Petropro-
represent the main reservoir interval, whereas the ducción eds., Quito, v. 2, p. 907–935.
Bourgois, J., A. Egüez., J. Butterlin, and P. De Wever, 1990,
middle Eocene clastics provide thin secondary reservoirs. Evolution géodynamique de la Cordillère Occidentale des
However, the alternation of periods of high sedimenta- Andes d’Equateur : la découverte de la formation éocène
tion rates (latest Paleocene and Lutetian) with periods of d’Apagua: Comptes Rendus à l’Académie des Sciences de
nondeposition and even emergence and erosion (early Paris, sér. II, v. 311, p. 173–180.
and late Eocene) (Figures 14, 15) probably disturbed Bristow, C. R., 1975, On the age of the Zapotal sands of
maturation of the organic matter. southwest Ecuador: Newsletter on Stratigraphy, Stuttgart,
Finally, the late Paleocene, early Eocene, and middle v. 4, p. 119–134.
Eocene tectonic phases recorded in the Santa Elena Bristow, C. R., 1976, The age of the Cayo Formation, Ecuador:
Peninsula have significance for exploration. The intense Newsletter on Stratigraphy, Stuttgart, v. 4, p. 169–173.
cleavage of some Cretaceous–Paleocene volcanic and Bristow, C. R., and R. Hoffstetter, 1977, Ecuador: Lexique
International de Stratigraphie, CNRS ed., Paris, v. Va2,
volcaniclastic rocks resulted in pronounced fracture 410 p.
porosity which has enhanced reservoir potential, such as Brown, C. B., and Baldry, R. A., 1925, On the Clay Pebble Bed
the small Santa Rosa–Petropolis oil field west of Santa of Ancón (Ecuador): Journal of the Geological Society,
Elena. Thermal perturbations and weak metamorphism London, v. 81, p. 454–460.
related to early tectonic events may have also modified Caldas, J., O. Palacios, V. Pecho, and C. Vela, 1980, Geología
and helped the organic maturation process. Although the de los cuadrángulos de Boyovar, Sechura, La Redonda,
geometry of the large-scale thrust planes cross cutting Punta La Negra, Lobos de Tierra, Las Salinas y Morrope:
the deformed pre-Lutetian beds is still virtually Boletin del Instituto de Geología, Minería y Metalurgía,
unknown, it undoubtedly controlled the migration and Lima, serie A, v. 32, 78 p.
trapping of hydrocarbons in the Santa Elena Peninsula. Canfield, R. W., 1966, Reporte geológico de la costa ecuato-
riana: Informe del Ministerio Industría y Comercio,
Finally, reverse faulting of probable late Eocene age that Asesoria Técnica de Petróleos, Quito, 150 p.
cross cut the older structures is responsible for east- to Carey, S., and H. Sigurdsson, 1982, A model of volcanogenic
southeast-dipping tectonic slices, within which small sedimentation in marginal basins, in B. P. Kokelaar and M.
reservoirs have been preserved. F. Howells, eds., Marginal basin geology: Geological
Society of London Special Publication 16, Oxford,
Blackwell Scientific Publications, p. 37–58.
Colman, J. A. R., 1970, Guidebook to the geology of the Santa
Elena Peninsula: Ecuadorian Geological and Geophysical
Acknowledgments This work is a contribution to IGCP 301 Society, Quito.
“Paleogene of South America.” We acknowledge Petroecuador Contreras, M., 1990, Estudio estratigráfico detallado de la
for permitting the publication of this study, and ORSTOM for Formación San Mateo en la localidad tipo, Manabí,
having supported field work. We are grateful to A. Alemán, G. Ecuador: Ing. dissertation, Escuela Superior Politécnica del
Laubacher, G. Mascle, and P. Roperch for having shared their Litoral, Guayaquil, 183 p.
630 Jaillard et al.

Cushman, J. A., and R. M. Stainforth, 1951, Tertiary Lebrat, M., F. Mégard, C. Dupuy, and J. Dostal, 1987,
foraminifera of Coastal Ecuador: part I, Eocene: Journal of Geochemistry and tectonic setting of pre-collision Creta-
Paleontology, v. 25, p. 129–164. ceous and Paleogene volcanic rocks of Ecuador: GSA
Daly, M. C., 1989, Correlations between Nazca/Farallón plate Bulletin, v. 99, p. 569–578.
kinematics and forearc basin evolution in Ecuador: Macharé, J., M. Sébrier, D. Huaman, and J.-L. Mercier, 1986,
Tectonics, v. 8, p. 769–790. Tectónica cenozoica de la margen continental peruana:
Egüez, A., and J. Bourgois, 1986, La formación Apagua: edad Boletín de la Sociedad Geológica del Perú, Lima, v. 76,
y posición estructural en la Cordillera occidental del p. 45–77.
Ecuador: Actas del IV Congreso Ecuatoriano de Ingeniería, Marchant, S., 1961, A photogeological analysis of the structure
Geología, Minería y Petróleo, Quito, v. I, p. 161–178. of the western Guayas province, Ecuador, with discussion
Egüez, H., 1985, Sedimentología y estratigrafía de la of the stratigraphy and Tablazo Formation, derived from
Formación San Mateo, Provincia de Manabí, Ecuador: Ing. surface mapping: Journal of the Geological Society,
dissertation, Escuela Superior Politécnica del Litoral, London, v. 117, p. 215–232.
Guayaquil, 111 p. Marchant, S., and C. D. G. Black, 1960, The nature of the Clay-
Faucher, B., R. Vernet, G. Bizon, J. J. Bizon, N. Grekoff, M. Lys, Pebble beds and associated rocks of south-west Ecuador:
and J. Sigal, 1971, Sedimentary formations in Ecuador; a Journal of the Geological Society, London, v. 115,
stratigraphic and micropaleontologic survey: Bureau p. 317–338.
d’Études Industrielles et de Coopération de l’Institut Marksteiner, R., and A. Alemán, 1991, Coastal Ecuador,
Français du Pétrole (BEICIP), 3 vol., 220 p. technical evaluation agreement: Unpublished internal
Faucher, B., and Savoyat, E., 1973, Esquisse géologique des report, Amoco Production Co. & Petroecuador, v. 1, 218 p.
Andes de l’Équateur: Revue de géographie physique et de Marocco, R., R. Baudino, and A. Lavenu, 1995, The intermon-
géologie dynamique, Paris, v. 15, p. 115–142. tane Neogene continental basins of the central Andes of
Feininger, T., and C. R. Bristow, 1980, Cretaceous and Ecuador and Peru: Sedimentologic, tectonic, and geody-
Paleogene history of coastal Ecuador: Geologische namic implications, in A. J. Tankard, R. Suarez, and H. J.
Rundschau, v. 69, p. 849–874. Welsink, Petroleum basins of South America: AAPG
Feininger, T., and M. K. Seguin, 1983, Simple Bouguer gravity Memoir 62, this volume.
anomaly field and the inferred crustal structure of conti- Mégard, F., 1987, Cordilleran and marginal Andes: a review
nental Ecuador: Geology, v. 11, p. 40–44. of Andean geology north of the Arica elbow (18° S), in J.
Gamber, J. H., G. W. Barker, J. A. Stein, J. L. Carney, A. F. W. H. Monger and J. Francheteau, eds., Circum-Pacific
Geen, A. F. Krebs, R. A. Salomon, and R. J. White, 1990, belts and evolution of the Pacific ocean basin: American
Biostratigraphic report on Coastal Ecuador: Unpublished Geophysical Union, Geodynamic series, v. 18, p. 71–95.
Technical Report, Amoco Production Co., Guayaquil, 65 p. Moreno, A., 1983, Estratigrafía detallada del Grupo Azúcar en
González, G., 1976, Bioestratigrafía del Eoceno en la región de los acantilados de Playas: Ing. dissertation, Escuela
Talara: Ing. dissertation, Universidad Nacional San Superior Politécnica del Litoral, Guayaquil, 182 p.
Agustin Arequipa, 225 p. Navarrete, E., 1986, Estudio micropaleontológico de la
Goossens, P. J., and W. I. Rose, 1973, Chemical composition Formación San Mateo en el corte de Puerto López-Salango,
and age determination of tholeiitic rocks in the Basic Creta- Manabí: Actas del IV Congreso Ecuatoriano de Ingeniería,
ceous Complex, Ecuador: GSA Bulletin, v. 84, p. 1043–1052. Geología, Minería y Petróleo, Quito, v. I, p. 111–122.
Goossens, P. J., W. I. Rose, and D. Flores, 1977, Geochemistry Nuñez del Arco, E., F. Dugas, and B. Labrousse, 1986,
of tholeiites of the Basic Igneous Complex of northwestern Contribución al conocimiento estratigráfico, sedimen-
South America: GSA Bulletin, v. 88, p. 1711–1720. tológico y tectónico de la región oriental de la Península
Hall, M. L, and J. Calle, 1982, Geochronologic control for the Santa Elena y parte Sur de la cuenca del Guayas (Ecuador)
main tectonic-magmatic events of Ecuador: Earth Science en base a 17 hojas geológicas escala 1/50.000°: Actas del III
Reviews, v. 18, p. 215–239. Congreso Ecuatoriano de Ingeniería, Geología, Minería y
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of Petróleo, Guayaquil, v. I.B., 33 p.
fluctuating sea levels since the Triassic: Science, v. 235, Olsson, A. A., 1931, Contributions to the Tertiary paleon-
p. 1156–1167. tology of northern Peru, part 4, the Peruvian Oligocene:
Henderson, W. G., 1979, Cretaceous to Eocene volcanic arc Bulletin of American Paleontology, v. 17, p. 100–264.
activity in the Andes of northern Ecuador: Journal of the Olsson, A. A., 1942, Tertiary deposits of north-western South
Geological Society of London, v. 136, p. 367–378. America and Panama: Proceedings of the 8th American
Jiménez, N., and E. Mostajo, 1990, Zonación de nanofósiles Sciences Congress, Washington, D.C., v. 4, p. 231–287.
calcáreos del Eoceno, Punta Ancón–Punta Mambra: Paredes, M., 1958, Terciario de La Brea y Pariñas y area de
Geociencia, Guayaquil, v. 3, p. 24–29. Lobitos: Ing. dissertation, Universidad Nacional San
Juteau, T., F. Mégard, L. Raharison, and H. Whitechurch, Agustin, Arequipa, 35 p.
1977, Les assemblages ophiolitiques de l’Occident équato- Pichler, H., and S. Aly, 1983, Neue K-Ar Alter plutonischer
rien: nature pétrographique et position structurale: Bulletin Gesteine in Ecuador: Zeitblatt der Deutschen Geologische
de la Société Géologique de France, v. 19, p. 1127–1132. Gesellschaft, Hannover, v. 134, p. 495–506.
Karig, D. E., and G. F. Moore, 1975, Tectonically controlled Reyes, L., and J. Caldas, 1987, Geología de los cuadrángulos
sedimentation in marginal basins. Earth and Planetary de Las Playas, La Tina, Las Lomas, Ayabaca, San Antonio,
Science Letters, v. 26, 233–238. Chulucanas, Morropon, Huancabamba, Olmos y
Lavenu, A., C. Noblet, M. Bonhomme, A. Egüez, F. Dugas, Pomahuaca: Boletín del Instituto de Geología Minería y
and G. Vivier, 1992, New K-Ar age dates of Neogene and Metalurgía, Lima, serie A, v. 39, 83 p.
Quaternary volcanic rocks from the Ecuadorian Andes: Romero, J., 1990, Estudio estratigráfico detallado de loa acanti-
implications for the relationship between sedimentation, lados de Machalilla, Provincia de Manabí: Ing. dissertation,
volcanism, and tectonics: Journal of South American Earth Escuela Superior Politécnica del Litoral, Guayaquil, 259 p.
Sciences, v. 5, p. 309–320.
Basin Development in an Accretionary Fore-Arc Setting, Southern Coastal Ecuador 631

Roperch, P., F. Mégard, C. Laj, T. Mourier, T. Clube, and C. Suess, E., Von Huene, R., and the Leg 112 Shipboard Scientific
Noblet, 1987, Rotated oceanic blocks in western Ecuador: Party, 1988, Introduction, objectives, and principal results,
Geophysical Research Letters, v. 14, p. 558–561. Leg 112, Peru continental margin, in E. Suess, R. Von
Santos, M., F. Ramírez, G. Alvarado, G. Guevara, and S. Huene et al., Proceedings of the Ocean Drilling Program,
Salgado, 1986a, La Formación Punta Blanca y su miembro Initial Reports, v. 112, p. 5–23.
San Mateo: Actas del IV Congreso Ecuatoriano de Inge- Thalmann, H. E., 1946, Micropaleontology of Upper Creta-
niería. Geologia, Minería y Petróleo, Quito, v. I, p. 49–60. ceous and Paleocene in Western Ecuador: AAPG Bulletin,
Santos, M., F. Ramírez, G. Alvarado, and S. Salgado, 1986b, v. 30, p. 337–347.
Las calizas del Eoceno medio del occidente ecuatoriano y Von Huene, R., and S. Lallemand, 1990, Tectonic erosion
su paleogeografía: Actas del IV Congreso Ecuatoriano de along the Japan and Peru convergent margins: GSA
Ingeniería, Geologia, Minería y Petróleo, Quito, v. I, Bulletin, v. 102, p. 704–720.
p. 79–90. Von Huene, R., and D. W. Scholl, 1991, Observations at
Santos, M., and F. Ramírez, 1986, La Formación Apagua, una convergent margins concerning sediment subduction,
nueva unidad eocénica en la cordillera occidental ecuato- subduction erosion, and the growth of continental crust:
riana: Actas del IV Congreso Ecuatoriano de Ingeniería, American Geophysical Union, Reviews of Geophysics, v.
Geologia, Minería y Petróleo, Quito, v. I, p. 179–190. 29, p. 279–316.
Sauer, W., 1965, Geología del Ecuador: Ministerio de Von Huene, R., L. D. Kulm, and J. Miller, 1985, Structure of
Educación ed., Quito, Ecuador, 383 p. the frontal part of the Andean convergent margin: Journal
Scholl, D. W., R. Von Huene, T. L. Vallier, and D. G. Howell, of Geophysical Research, v. 90, p. 5429–5442.
1980, Sedimentary masses and concepts about tectonic Wallrabe-Adams, H.-J., 1990, Petrology and geotectonic devel-
processes at underthrust ocean margins: Geology, v. 8, opment of the western Ecuadorian Andes: the Basic
p. 564–568. Igneous Complex: Tectonophysics, v. 185, p. 163–182.
Schulman, N., A. Flexer, and E. Washkal, 1965, Geology and
groundwater possibilities of central Manabi, Ecuador:
Ministry of Foreign Affairs, Department of International
Cooperation, Israel.
Séranne, M., 1987, Informe geológico sobre la evolución
tectónica y sedimentaria de la cuenca Talara: Unpublished
Authors’ Mailing Addresses
Internal Report of the Instituto Frances de Estudios Étienne Jaillard
Andinos and Petróleos del Perú, Lima, 73 p. ORSTOM
Shepherd, G. L., and R. Moberly, 1981, Coastal structure of the 213, rue La Fayette
continental margin, northwest Peru and southwest
Ecuador: GSA Memoir, n. 154, p. 351–391.
75480 Paris Cédex 10
Sheppard, G., 1937, The geology of southwestern Ecuador: France
London, Billing and Sons, 275 p.
Sigal, J., 1969, Quelques acquisitions récentes concernant la Martha Ordoñez
chrono-stratigraphie des formations sédimentaires de Stalin Benitez
l’Équateur: Revista Española de Micropaleontología, v. 1, Gerardo Berrones
p. 205–236. Nelson Jiménez
Sinclair, J. H., and C. P. Berkey, 1923, Cherts and igneous Galo Montenegro
rocks of the Santa Elena oil-field, Ecuador: Transactions, Italo Zambrano
American Institute of Mining and Metallurgical Engineers, Petroproducción
Canadian meeting, Montreal, v. 69, p. 79–95.
Small, J., 1962, Stratigraphy of southwest Ecuador and Ancón
km 6.5 via a la Costa, casilla 10829
oilfield studies: Ph.D. dissertation, Massachusetts Institute Guayaquil
of Technology, Massachusetts, 185 p. Ecuador
Eastern Cordillera of Colombia: Jurassic–Neogene
Crustal Evolution

D. Roeder R. L. Chamberlain
Institut für Lithosphärenforschung Blue Eagle Exploration Inc.
Justus Liebig–Universität Salisbury, North Carolina, U.S.A.
Giessen, Germany

Abstract

T he Eastern Cordillera of Colombia is an east-vergent Dahlstromian fold and thrust belt of Neogene age
involving 5–10 km of upper crust, 3 km of synrift fill of an early Atlantic embayment, and 11 km of
Cretaceous thermal sag basin fill. The Mesozoic rift event created an estimated 67–100 km of extension in a 200-
km-wide field with an average extensional strain (β) of 2. The fold and thrust belt contains 170 km of imbricate
thrust overlap at its front and at higher thrusts and 60 km of west-vergent back-thrusting at its west slope and
at its crest affected by incipient plateau collapse tectonics. The Oriente foreland load flexure suggests a rigid
lithosphere and more than 100 km of intracontinental subduction.
We have assembled and numerically processed public domain data about the surface geology, Mesozoic
basin history, and subsurface structure in the Oriente foothills. Andean crustal data and northeastern Pacific
plate tectonic data suggest a Miocene age change from low-dip subduction to Recent steep-dip subduction.
Therefore, a polyphase history of foreland upthrusts redeforming the fold and thrust belt is unlikely. However,
upper crustal detachment and foreland upthrusting prior to Cordilleran thrusting is possible.

Resumen

L a Cordillera Oriental de Colombia es una faja fallada y plegada con vergencia oriental, de estilo similar
al definido por Dahlstrom y edad neógena. Involucra 5 a 10 km de corteza superior, 3 km de relleno
sintectónico depositado en el inicio de una bahia atlántica, y 11 km de sedimentitas cretácicas relacionadas a
subsidencia termal. La extensión mesozoica está estimada en 67 a 100 km en una franja de 200 km de ancho,
con una deformación promedio (β) de 2. La faja plegada y fallada posee 170 km de acortamiento en los corrim-
ientos imbricados en su frente oriental y corrimientos someros, y 60 km en los retrocorrimientos de vergencia
occidental en su pendiente oeste. La porción crestal está afectada por una tectónica extensional incipiente. El
flexuramiento por carga del antepais oriental sugiere una litósfera rígida y mas de 100 km de subducción intra-
continental.
Fueron reunidos y procesados numéricamente datos de dominio público acerca de la geologia de super-
ficie, desarrollo de la cuenca mesozoica y estructuras en el subsuelo pedemontano oriental. El análisis de los
datos relacionados a la corteza andina y a la tectónica en el noreste de la placa pacífica sugieren un cambio en
la inclinación de la subducción, de baja durante el Mioceno a mas empinada en épocas recientes. De acuerdo a
lo expuesto, es poco probable un desarrollo polifásico de corrimientos redeformando la faja plegada y fallada.
Por lo tanto, existirían despegues en la corteza superior y deformación del antepaís anteriores al fallamiento
cordillerano.

Roeder, D., and R. L. Chamberlain, 1995, Eastern Cordillera of Colombia: Jurassic– 633
Neogene crustal evolution, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum
basins of South America: AAPG Memoir 62, p. 633–645.
634 Roeder and Chamberlain

INTRODUCTION
In central Colombia (Figure 1), the Eastern Cordillera 500 km C
rises out of the Atlantic initial embayment, which is the
depositional site of one of the largest hydrocarbon source
rock bodies on earth (Shirley, 1992). By restoring the fill M
P
of this rifted embayment and by involving the crust in
balancing efforts along an Andean cross section, it may
be possible to decide among three different geologic
(3)
interpretations. (1)
Our preferred interpretation involves Neogene upper
crustal and supracrustal low-angle thrust imbrication (4)
above a cratonic lithosphere. The allochthonous crust B
exhibits at least 170 km of crustal overlap on an east- (2)
vergent (frontal) imbricate thrust system and involves
30–60 km of internal compression and back-thrusting. It
shows a crustal volume deficiency despite a 40-km Q
(+8 km) deep Moho root. We see the volume deficiency
as inherited from the Mesozoic Atlantic-type rifting and
not as the product of a late Andean episode of upper Figure 1—Location map of northwestern South America,
including Colombia and parts of Panama, Venezuela,
crustal foreland upthrusting. However, early Andean Ecuador, and Peru. Light gray, Guyana shield; dark gray,
foreland upthrusting at a future site of the present Andean orogenic elements, with thrust faults at edges.
Eastern Cordillera is possible. Indexed cross sections are from (1) Colletta et al. (1990),
These tectonic results are consistent with Andean (2) Valderrama (1982), and (3) Meissnar et al. (1976);
subduction of the steep or steepening type during the (4) Peru-Chile trench. Cities: P, Panama; M, Maracaibo; C,
Neogene, with Neogene volcanism, and with the Caracas; B, Bogota; Q, Quito.
Neogene trend of Andean crustal thickening crossing the
Mesozoic Atlantic embayment. In this paper, we use the
term Dahlstromian to refer to a structural geometry and consistent with major wrench faulting and can be
tectonic rheology developed for subsurface exploration presented without the wrench faults. However, our
of fold and thrust belts (e.g., Bally et al., 1966; Dahlstrom, results should be modified in the future to incorporate
1969), critically reviewed by Ramsay and Huber (1987) the effects of wrench faulting.
and formally named by Roeder (1991). In Figure 2, we juxtapose three major competing and
The present study adds a third interpretation by using in part speculative interpretations of the standard cross
the database and selected parts of the geologic theory section. All three versions are valid geologic achieve-
used in two competing and complementing published ments, and their review, even at their incompletely docu-
cross sections (Colletta et al., 1990; Dengo and Covey, mented state, can help in guiding the major investments
1993). It is also based on a field reconnaissance in 1982 needed in sub-Andean hydrocarbon exploration. We
with H. Doust (The Hague) and J. Kirkpatrick (Denver) have standardized and simplified the graphic appear-
and on an introduction to Andean structure in 1982 by R. ance of the three versions, and to each section we have
Butler (Columbia, SC). added a best guess of the present crustal thickness.
We have extrapolated the Moho position 650 km to
the north from the trans-Andean geophysical profile of
THREE TECTONIC MODELS Colombia and Ecuador (Case et al., 1973) by using trends
shown on the Bouguer gravity map of Colombia
As shown in many cross sections (De Cizancourt, (Bermudez et al., 1985) and using an assumed confor-
1933; Campbell and Bürgl, 1965; Campbell, 1974; mity between a northward-decreasing Bouguer anomaly
Estrada, 1982; Butler, 1983), the Eastern Cordillera of and a northward size reduction of the Moho root (Ocola
Colombia is a bivergent fan of thrusts toward the Middle et al., 1975; Meissnar et al., 1976). The flank dips of the
Magdalena and Oriente lateral foredeeps. It coincides Moho root are constrained by flexural load models of the
with a 9–12-km-thick Jurassic and Cretaceous depo- Oriente and Middle Magdalena foredeep basins.
center. A standard cross section 200 km long between the In all three sections in Figure 2, the 36-km-deep Moho
Magdalena river and the Oriente lowlands near Aguazul root implies an excess crustal depth of 19 km and an
and Cusiana (Figure 2) crosses this fold and thrust fan at excess crustal area of 1500 km2. As in other orogens, the
a site covered by quadrangle mapping (Ulloa and excess area can be explained by magmatic addition
Rodriguez, 1978; Renzoni et al., 1983). (James, 1971) or by crustal overlap (Giese et al., 1982;
Mapped in great detail and over long distances, strike- Mattauer, 1986). The choice between these alternative
slip faults are a major constituent of Andean tectonics in explanations is not apparent from direct local observa-
the Eastern Cordillera of Colombia. In the present tion. However, in the three parts of Figure 2, the juxtapo-
context, however, wrench faults are virtually ignored. sition of upper crustal silhouettes and the Moho root
The arguments developed in the present paper are design shows varying degrees of conformity.
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution 635

(a) COLLETTA ET AL. (1990)

50 KM
(b) DENGO AND COVEY (1993)

50 KM

(c) ROEDER AND CHAMBERLAIN (1994)

50 KM
Figure 2—Three versions of a standard cross section of the Eastern Cordillera, central Colombia. Black, Cretaceous–
Paleogene sedimentary rocks; white, Jurassic; gray, upper crust and predeformed Paleozoic metasedimentary rocks. An
estimated best-fit curve of the Moho seismic velocity discontinuity has been added. (a) Redrawn after Colletta et al. (1990),
extended downward to include the base of the brittle upper crust and estimated Moho position. (b) Redrawn after Dengo and
Covey (1993), with the base of the brittle upper crust and estimated Moho position added. (c) Prepared for present paper,
obtained by applying a mobilistic interpretation of Neogene tectonics to the cross section by Dengo and Covey (1993).

The eastward-sloping Moho segment beneath the initiated as rift flanks and extensional normal faults.
Magdalena Valley is a result of the projection of the Between compressionally reactivated faults or upthrusts,
Moho root along strike. Its connection with the second upper crustal panels dip toward the the orogenic center
Moho deep beneath the Central Cordillera is not docu- more steeply than the asumed best-fit Moho. Disre-
mented. This aspect is a point in favor of the crustal garding unmapped strain, the shortening by reverse
thrust model (Dengo and Covey, 1993). faulting shown in this cross section is 29 km.
Integrated over a width of 130 km, this bulk strain is
First Tectonic Model: Inverted Rift Basin equivalent to a negative elongation of 18%, and it would
thicken the 10-km-thick undeformed lower crust to 12
and Upthrusts km. Applied to the total crust and disregarding erosion,
Figure 2A illustrates the interpretation by Colletta et the bulk strain represents an excess cross-sectional area
al. (1990) in a redrawn form. Colletta et al. (1990) have of 650 km2, about half of the excess area of the Moho
explained the Eastern Cordilleran bivergent fan as a root. This mismatch is purely geometrical because it is
compressional inversion of an extensional basin complex poorly constrained and because it disregards important
and its rift fill. Fault blocks of pre-Jurassic rocks mapped geologic aspects of rift formation and inversion. The
by Colletta et al. (1990) are shown to 12.5 km below mismatch, therefore, is only a weak argument against the
basement top. To facilitate the comparison of all three tectonic model of Colletta et al. (1990).
interpretations, we have extended the faults downward
at a constant angle as mapped. At a geotherm of Second Tectonic Model: Low-Angle
24°C/km, the base of the fault blocks represents the Thrusting Followed by Foreland Basement
brittle–ductile transition just prior to fault displacement. Upthrusts
Dips and displacements of the faults reflect an upthrust
style (Harding and Lowell, 1979; Bally et al., 1985) of low Figure 2B is a new display of the interpretive cross
strain, inverting steeply dipping rift flanks and strike-slip section of Dengo and Covey (1993). This is the most
faults. As is typical for inverted rift fill (Bally et al., 1985), advanced and best documented of the three interpreta-
faults are assumed or observed to steepen downward. tions shown. Its authors propose a Neogene succession
Many of the faults shown are assumed to have of two groups of events affecting a Mesozoic passive
636 Roeder and Chamberlain

extensional margin, its sedimentary prism, and its REFLECTION SEISMIC DATA: ORIENTE
oceanward-thinning crust. Andean subduction transmits
part of its compression into an eastward-vergent
FOOTHILLS
supracrustal low-angle thrust fault. This thrust propa-
gates through the Mesozoic sedimentary strata as a Figure 3 is based on seismic data suggesting that the
complex bundle of ramping, back-thrusting, and eastern mountain front of the Cordillera Oriental
upward-soling Dahlstromian thrust faults (Suppe, 1983; (Colombia) is part of a fold and thrust belt and not part
Roeder, 1991). After an estimated 105 km of emplace- of a basement upthrust province. The figure is a repro-
ment (determined by line balancing and snip restoring), portioning, or squash plot, of a reflection seismic profile
the upper crustal thrust system is overprinted by steeply shot that has been processed and interpreted by industry
dipping basement-involved foreland upthrusts. These (Triton Colombia proprietary report). The seismic line
late Andean structures bring up the basement, show crosses the Cusiana trend in the footwall of the Yopal
strike slip, and are the sites of observed and historical thrust and is located parallel to the cross section (Colletta
seismicity. et al., 1990) and offset at an unspecified distance.
The Oriente foredeep is explained only in part by The seismic line shows an undisturbed foreland with
upper crustal overlap load, but most of the steep west dip about 4 sec (two-way time) of conformable events from
in the foothills segment of the upper crust is left unex- Tithonian–Miocene basin fill, perhaps unconformably
plained. The mountainward dip of the Magdalena basin overlying Paleozoic rocks below 4 sec. In the north-
floor is explained by Dahlstromian tectonics in the upper western third, a northwest-dipping series overlies and
crust. The basin floor is the bedding system of a hanging truncates an antiformally warped middle unit. The
wall ramp transported on top of a footwall flat. This contact intersects the edge of the line at 3 sec, and it is
crustal doubling may form a transition to the thick crust clearly a thrust fault in a flat-ramp position. Its surface
beneath the Central Cordillera (not shown in Figure 2). trace has been mapped as the Yopal thrust (Hebrard,
Kinematic or orogenic successions of transcrustal low- 1985; Colletta et al., 1990). At its base, the antiformally
angle thrusting succeeded by upper crustal foreland warped middle unit also truncates reflections belonging
detachment are common not only in the Andes (Jordan to the foreland series. This contact can be interpreted as a
and Allmendinger, 1986) and the North American ramp-on-ramp thrust fault.
Cordillera (Miller et al., 1992) but also in the Alps The ramp top of this thrust fault does not visibly
(Trümpy, 1980) and other collision belts (Dewey et al., break the topographic surface. Most likely it merges with
1986). This would support the kinematic interpretation a blind top detachment at about 1.5 sec. Figure 3 does not
by Dengo and Covey (1993). However, we suggest a show, but would be consistent with, a coherent foreland
different interpretation of this succession of styles. reflector at the base of the detached series. Based on this
The interpretation by Dengo and Covey (1993) is interpretation, the Oriente foothills show a minimum of 8
consistent with a Talwani-type Bouguer gravity model km of thrust displacement and a structural style diag-
closely fitting the available data. However, there is poor nostic for detached fronts of fold and thrust belts (e.g.,
conformity between the upper crustal blocks and the Mitra, 1986, 1992; Bally et al., 1985; Sobornov, 1991; Vann
assumed best-fit Moho. et al., 1986).

Third Tectonic Model: Crustal Thrusts RIFT TECTONICS


Prograde Eastward and Load a Cratonic
Foreland Rift systems worldwide (Williams et al., 1989;
Letouzey, 1990; Ziegler, 1990) confirm the geologic
Figure 2C shows our proposed interpretation. It is validity of the thermal and isostasy-based model of two-
closely based on the surface geology as presented by phased rift subsidence known as the McKenzie model
Dengo and Covey (1993), and it has been developed with (McKenzie, 1978; Le Pichon and Sibouet, 1981). The
additional data, procedures, and speculation to be tectonics of rift systems are an essential support for our
discussed in the following pages. palinspastic estimates and hence for our tectonic model.
We have assumed (Hatcher and Hooper, 1992) that Thus, rift tectonics are summarized in Figure 4 and as
upper crust can be involved in low-angle thrusts, and we follows:
have modified some key details in the cross section by
Dengo and Covey (1993). We have also fitted a re- 1. A synrift phase of extension leads to isostasy-
designed thrust architecture into the space outlined by driven subsidence, intracrustal or translithospheric
the load-deflected crust of the Oriente basin. We have low-angle extensional detachments, and graben
assumed that the tilt of the Magdalena basin is due to an structures with diagnostic synrift fill of sedimen-
elastic line load flexure and have achieved conformity tary and/or magmatic rocks. The amount of synrift
between the thrust architecture and the assumed best- subsidence depends on the extensional strain (β),
fitting Moho. We interpret the back-thrusting in the top the original thickness of the rifted sialic crust, and
part of the thrust stack as synkinematic wedge thick- the density of the rift fill. At infinite (β) extension,
ening during uphill emplacement along a density such as at an ocean–continent interface with
gradient between crust and mantle. normally thick sialic crust, an average synrift fill
0

TM
TM 5 KM
L 1

L C
C M L 2
M CB C
CB G M
G CB 3
G
CS
4

PAL
BSMT
5

Figure 3—Reflection seismic profile across the east front of the Eastern Cordillera near the oil field of Cusiana (Colombia). Hand-made line tracing after unpublished
data acquired in 1982 by Triton Colombia, Inc., displaying a predrilling geologic interpretation by Anonymous of Triton, Inc. Vertical scale is in seconds two-way time.
Shaded lines with letters are stratigraphically identified seismic reflectors. TM, Guyabo (Miocene); L, near top of Leon (Oligocene); C, near top of Carbonera
(Oligocene); M, near top of Mirador (Eocene); CB, near top of Cuervos-Barco (Paleocene); G, near top of Guadalupe (Upper Cretaceous); CS, Upper Cretaceous; PAL,
perhaps Paleozoics; BSMT, near top of crystalline basement.
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution
637
638 Roeder and Chamberlain

POSTRIFT, THERMAL SUBSIDENCE, LOAD-DEFLECTED FILL

SYNRIFT ZONE SYNRIFT ZONE


OF LOW SHEAR AND HALF-GRABENS OF HIGH SHEAR AND TECTONIC DENUDATION

PRERIFT LITHOSPHERE PRERIFT LITHOSPHERE


DETACHED HANGING WALL FLANK LOAD-DEFLECTED FOOTWALL FLANK
50 KM H + V

Figure 4—Schematic cross section of a site of McKenzie-type extension or longhorn basin, developed from a calculated load
flexure and a snip restored transcrustal low-angle shear. Black, postrift sedimentary rocks; gray, crust. See discussion in
the text.

thickness of 8.5 km is predicted by the McKenzie subsidence curves. However, stratigraphic control is
model. At an intracratonic rift with 1.5 (β) exten- needed to identify synrift and postrift deposits.
sion, a 2-km-thick synrift fill will accumulate.
2. Restoration of the rift-induced thermal anomaly is Longhorn Basin Geometry
driven by the amount of thermal dispersivity that
is specific for silica (Oxburgh and Turcotte, 1974; Figure 4 shows an important two-dimensional aspect
Turcotte and Schubert, 1982). It causes subsidence of the strictly one-dimensional and Airy-type McKenzie
at an exponentially decaying rate with a time model as discussed. Subsidence during the rift and
constant of 62.8 Ma. The amount of subsidence postrift phases regionally affects slabs of lithosphere by
depends on the time point and span of thermal flexurally loading them. The effect of flexural loading
subsidence observed and on the density of the generates a basin shape informally called longhorn or
sediment fill, which typically consists of a passive steer’s head basins.
margin series. The stratigraphic record shows the Thermal subsidence takes place during tectonic quies-
exponentially decaying subsidence rate. At a 120- cence. It acts as a line load (Heiskanen and Meinesz,
Ma ocean–continent interface, 6 km of thermal 1958), affecting not only the rift site but also the quiescent
subsidence fill (or passive margin fill) is predicted forelands of the rift with a predictable exponential decay
by the McKenzie model. in the dimension of distance away from the rift. Thermal
subsidence basins or longhorn basins are sag shaped
Polyphase Rifting with a depocenter over the rift site and a zero edge at a
distance proportional to the elastic wavelength of
Complex plate configurations in young oceans cratonic lithosphere. This distance is commonly between
commonly generate more than one rift tectonic event. 100 and 500 km from the rift site (Watts and Ryan, 1976).
The subsidence effects of successive events will overlap We interpret the Tithonian–Paleocene succession of the
and generate errors in predicting thicknesses of actually Eastern Cordillera to be a pre-Andean longhorn basin.
deposited stratigraphic units. This may explain why the Figure 5 is a segment of the restored cross section by
thickness of the Eastern Cordilleran Cretaceous substan- Colletta et al. (1990) with a vertical exaggeration of 400%.
tially exceeds the thickness predicted by the McKenzie A silhouette at the top of Figure 5 shows the section as
model. To apply the McKenzie model appropriately to published (Colletta et al., 1990). Our selected segment
the Cretaceous depocenter of the Eastern Cordillera of covers the area between the Oriente foothills and the
Colombia, it is important to decide if any of the discor- Comichoque thrust near Tunja. It includes the strati-
dant stratigraphic units are rift fill or overlapping graphic succession of Tithonian–Paleocene/Maas-
sediment aprons deposited during phases of thermal trichtian age, with local pockets of Jurassic strata. Colletta
decay. This decision will be based more on stratigraphic et al. (1990) have interpreted the thickness variations as
analysis and less on precise numerical applications. caused by syn- or predepositional block faults.
Younger synrift tectonics may dissect the passive Our reinterpretation of the standard cross section of
margin apron of a slightly older thermal subsidence the Eastern Cordillera is based mainly on a different
event. In this setting, the second-phase synrift fill may interpretation of the rift tectonics. The Jurassic deposits in
have a postrift lithology. Polyphase rifted passive the Magdalena valley area are described as fluvial
margins may accumulate a multitude of the predicted gravels, wadi fill, and rhyolitic volcanic rocks (e.g.,
postrift fill. Extensional passive margin successions Butler, 1983; Arango et al., 1976; Renzoni et al., 1983). As
much thicker than 6 km are common. With enough shown in Figure 6, we interpret these deposits as synrift
control of thickness and ages, their stratigraphic record fill. We explain the Tithonian–Maastrichtian succession
may show the diagnostic pattern of stacked cuspate as shallow water to subtidal and open marine deposits
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution 639

deflection. This approach suggests that the eastern part


COCUY BASIN - CRETACEOUS FILL
of the Eastern Cordillera is underlain by three major low-
angle thrust faults with 44, 18, and 45 km of minimum
(a) TO SCALE thrust overlap. Although this method is inadequate for
precise conclusions, it is consistent with the reflection
seismic data (Figure 3). It confirms the conclusion that
the eastern front of the Eastern Cordillera of central
(b)
VERT EXAG 400 %
Colombia has been involved in major low-angle
thrusting. We now incorporate crustal data into this rein-
20 KM
terpretation and arrive at additional conclusions about
the structural style.

Figure 5—Stratigraphic cross section of the Cretaceous in


the Eastern Cordillera, Colombia. (A) Redrawn at 100%
(silhouette) and (B) at 400% vertical exaggeration, shaded CRUSTAL CONFIGURATION
with unidentified stratigraphic boundaries (after Colletta et Surface geology and Cretaceous basin evolution
al., 1990). Sites of abrupt changes in thickness are inter-
preted as rift walls affecting synrift deposits (Colletta et al.,
imply an orogenic shortening of 140–170 km across the
1990) or as thrust faults affecting a postrift succession Eastern Cordillera. This bulk strain should be expressed
(present paper). in the shape of the crust, even if there is no seismic record
of Moho stacking. The crustal structure of the northern
Andes is only vaguely known (Ocola et al., 1975;
Meissnar et al., 1976; Bermudez et al., 1985). The effects of
(e.g., Campbell and Bürgl, 1965; Arango et al., 1976; Tithonian rifting on the crustal base may contribute to
Butler, 1983). We therefore interpret the Tithonian–Maas- additional vagueness. Our flexural load modeling is
trichtian stratigraphy as the stacked and telescoped consistent with a 35–40-km-deep Moho root.
postrift (or thermal subsidence) part of a longhorn basin.
The finite thickness of the Tithonian–Maastrichtian
succession is assumed equivalent to the sediment-loaded Flexural Load Models
thermal subsidence, with an error introduced by The shape of the Andean crust and the amount of
ignoring variations in water depth of deposition. thrust overlap can be constrained if the depth of the
With a definition of its elastic load deflection, we can Moho root and the original crustal thickness are known.
quantitatively restore the thrust telescoping of Andean In the study area, Moho depth and original crustal
age. The deflection depends on the rigidity of the lithos- thickness are unknown, but estimates can be made. The
phere and is expressed in its elastic half-wavelength, constraining argument comes from tracing the elastic
which is a distance across strike of the deflection. The deflection of the crust under the composite orogenic load
longer the half-wavelength, the greater the implied (Turcotte and Schubert, 1982; Royden and Karner, 1984).
paleogeographic and restorable distances. The geometric aspects of the elastic deflection can be
For the Andean foredeep load, a half-wavelength of used in constructing structural cross sections.
400 km is constrained between the Moho root near Tunja Figure 7a is a simplified cross section of the
and the basement outcrops near the Guaviare River Oriente–Rio Meta basin (Valderrama, 1982). It includes
(Arango et al., 1976). For the longhorn basin, subsurface some well control, some intrabasinal tectonics (not
data in the closer foredeep realm (Valderrama, 1982) discussed here), and a poorly controlled zero edge of
suggest a telescoped half-wavelength of 200–300 km. The deflection and/or sedimentation near the Rio Guaviare
elastic rigidity of the South American lithosphere (labeled B). This cross section (Valderrama, 1982)
probably did not change significantly between the suggests a bimodality of crustal curvature, amplified
Tithonian and Neogene. However, we have based our here by a vertical exaggeration of 400%. At distances
restoration on a conservative flexural half-wavelength of greater than 100 km southeast of the orogenic front, the
230 km because it is probably a composite of several Llanos basin floor is nearly flat. Closer to the orogenic
rifting events. This half-wavelength is an estimate at the front, the basin floor appears more strongly flexed and it
low end of the geodynamic range, not a documented attains a mountainward slope of 6°. Part or all of this
number. increased flexure may be due to an erroneous accommo-
dation of telescoped thickness data. However, by
rigorous emulation of the basin shape by Valderrama
PALINSPASTIC RESTORATION (1982) and by honoring the data in well 2 (La Gloria), we
can show that the bimodal curvature is not an artifact.
For our restoration of the Tithonian–Maastrichtian Figure 7b shows one successful and several unsuc-
depocenter (Figure 6), we began by plotting out a 230- cessful flexure emulations. The left (northwestern) end
km-long, 9.3-km-deep line-loaded elastic flexure. We bracket of the diagram represents the deflected depth of
then arranged the segments identified as fault blocks in the Moho root. The top of the vertical bar in the center of
the inversion model (Colletta et al., 1990) by sliding them the diagram represents the base of the Cretaceous
along the flexure so that their thickness would fit the implied by well 2 (La Gloria).
640 Roeder and Chamberlain

A B C D E
55 KM 30 KM 18 KM 45 KM 8 KM

V. E. = 400 %

50 KM (H, NOT V)

A: PESCA THRUST CRETACEOUS FILL REINTERPRETED


B: SANTA BARBARA THRUST AS LINE-LOADING POSTRIFT SERIES
C: GUAICARAMO THRUST AFTER COLLETTA ET AL. (1990)
D: YOPAL THRUST

Figure 6—Stratigraphic cross section of Eastern Cordilleran Cretaceous data in Figure 5, from Cocuy basin, Colombia. Data
are palinspastically arranged along a numerically defined line-loaded flexure. The horizontal distances inserted between the
columnar and shaded segments are interpreted as distances lost during Andean fold and thrust tectonics.

(a)
1 2 3 4 5 A B

5 KM
100 KM
(b)

EMULATION WITH ELASTIC LOAD FLEXURE


SOLID: COMPOSITE BEAM, DASHED: SIMPLE BEAMS

Figure 7—Two cross sections of eastern Colombian Llanos basin and adjacent Eastern Cordillera. (a) Structural cross
section with four times vertical exaggeration (redrawn after Valderrama, 1982). Gray, Tertiary foredeep; black, Cretaceous
postrift; white, Jurassic and Paleozoic basement; A, possible site of zero deflection near Rio Guaviare; B, possible site near
mapped edge of Guyana shield. Well 2 is La Gloria. Internal tectonics of Valderrama (1982) are not shown. (b) Simulation of
basin structure from Valderrama (1982) with line-loaded flexural beams having different rigidities and maximum deflections.
See discussion in text.

A successful model is shown in two black lines. It Structural Implication of the Flexural Load
honors the shallow flat Llanos basin floor and assumes Model
an edge of zero deflection near point A in the shallow
and poorly controlled basin. The successful model also Geologic reality is certainly more complex than our
honors the Cretaceous base of well 2, as well as a crustal numerical emulation of load flexure, but we have used
deflection of 12.5 km near Tunja in the Eastern the flexure model in building the structural cross section.
Cordillera. We have achieved the successful emulation Differentiating the flexural load equations suggests
by adding two flexures. Combining several flexures is numerically that the foreland slope beneath the Eastern
geologically admissible as a form of polyphase deforma- Cordillera dips mountainward at 4.2°–4.5°. Following
tion because of the viscous aspects of the physical model Lyon-Caen et al. (1985), the implied rigidity suggests an
(Karig et al., 1976). Andean load overlap on the order of 100 km.
The first of the added flexures consists of a long wave The long-wave flexure reflects a cratonic lithosphere
flexure of a rigidity D = 1.7 × 1024 Nm. It is equivalent to with a small end load, such as a quiescent passive
a flexural length of 400 km, and it suggests a lithospheric margin succession. The flexure does not show the
elastic thickness of 50 km. The second of the added presence of the Andes or a major crustal subduction.
flexures is a short-wave flexure of 2.6 × 1023 Nm, which Maps by Valderrama (1982) show a northeastward and
is equivalent to a flexural length of 250 km and an elastic divergent strike of the long-wave deflection.
thickness of 26 km. All other possible flexure parameters The short-wave flexure is similar to the flexure
(including two of them in Figure 7b shown by dashed implied by the postrift succession of the longhorn basin.
lines) lead to curvatures of the foreland basement top Its implied crustal thickness is that of crust directly
which miss one or two of the required three control underlain by asthenosphere. This setting is equally
points. possible in extensional and compressional regimes (Bird,
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution 641

TUNJA - 1 UNETE - 1 MIRADOR - 1 A LA GLORIA

50 KM

Figure 8—Structural cross section of allochthonous parts of the southeastern half of the Eastern Cordillera, located at the
site of the cross section of Colletta et al. (1990) (Figure 2A). Black, Cretaceous–Paleogene; white, Upper Jurassic and
contents of buried Bogota basin; gray, Paleozoic rocks and upper crust. Cross section follows interpretation by Dengo and
Covey (1993) but admits some modifications.

1979; Brown et al., 1992). As in other passive margin imbricated beneath a roof thrust sheet composed of
edges, it could have survived the Tertiary and Andean Jurassic rocks above crystalline or Paleozoic rocks.
load deflection. However, the maps (Valderrama, 1982) The foothills are shown with the bed-parallel refolded
show that its strike is Andean. Therefore, either its origin detachment by Dengo and Covey (1993) at the base of
is Andean or Neogene, or the Andean orogenic belt the Neogene section involved in the Zapatosa syncline or
locally follows the trace of a Mesozoic lithospheric scar. equivalents. In the cross section, this thrust is shown to
merge with a frontal thrust that has emplaced basement
Eastern Cordillera: Present Structure or Paleozoic rocks over Cretaceous. From east to west,
the foreland succession below the frontal thrust contains
Figure 8 shows a portion of our reinterpretation of the the Cusiana structure or equivalent, an emergent group
cross section by Dengo and Covey (1993). Critical details of slabs forming salient mountain ranges and outliers,
of it are based on the arguments developed in the and a large buried thrust sheet with thick Cretaceous.
foregoing pages. This slice is only implied by the shape of the Zapatosa
At the assumed crustal thickness of 18 km (not shown syncline, and it is a structural petroleum prospect.
in Figure 8), a simple doubling with modest thrust The shape of the Bogota basin has been assumed from
overlap could explain the proposed best-fit Moho root. Dengo and Covey (1993), but we have added a west-
The supracrustal imbrication of 175 km suggested by the ward tilt. Its fill is shown with the Jurassic signature, but
longhorn model fits the flexural space only if the crust it could also consist of Cretaceous, as suggested by
has been severely thinned during the Jurassic extension. Dengo and Covey (1993).
To satisfy the gravity field (Dengo and Covey, 1993), we
assume that imbrication has concentrated on the upper Andean Geodynamics
crust and on detachment between upper crust and
mantle (Bird, 1979; Mattauer, 1986; Isacks, 1988). A complete geodynamic interpretation of the Andes
The concepts of thrust wedge dynamics (Davis et al., of Colombia would explain the amount of crustal and
1983) suggest that the Moho overlap may be located at sedimentary imbrication, the origin of this compression,
the base of the imbricate stack and that it ties in with the and its timing relative to plate tectonic events. Such an
Foothills detachment at the base of the Mesozoic interpretation is not possible with present data and
sediments (Figure 2). The footwall of this Moho overlap understanding. However, the speculative and unusual
is the load-deflected Llanos lithosphere. The hanging subsurface interpretation in the present paper may be
wall of the Moho overlap is a complex imbricate stack served by a discussion of its geodynamic significance.
that evolved out of the jumbled and block-faulted center This discussion may also help in exploring other Andean
of the former longhorn basin. segments and other foreland fold and thrust belts.
The cross section in Figures 2 and 8 shows three In our view, the prethrusting distance across the
crustal slices of the C type (Hatcher and Hooper, 1992) Eastern Cordillera is constrained by the restored width of
which may emerge at the surface along strike to the the Tithonian–Paleogene longhorn basin. This implies
north and to the south. Only the top slice shows signifi- large overlap on major thrust faults. The commonly
cant back-thrusting. The imbricate structure of upper exposed pre-Mesozoic rocks suggest that basement or
crust, Paleozoic, and Mesozoic sedimentary strata upper crust and Paleozoic tectonite are largely involved
suggests that the sediment panels between the C-type in the thrust slices.
slices are westward dipping. This generates more thrust
overlap and deeper pockets of sediments. This also Estimated Strain of Tithonian Rift
generates additional space beneath the Arabuco
anticline. We are therefore showing this structure as a To illustrate the postrift pre-Andean configuration of
duplex associated with the crystalline front as commonly the standard cross section, we have prepared a vaguely
realized in orogenic belts (Roeder, 1989). In the assumed line-balanced snip restoration (as described by Roeder,
Arabuco duplex, we have drawn the Cretaceous as 1991), shown in Figure 9. This restoration suggests that
642 Roeder and Chamberlain

X' 50 KM

50 KM

Figure 9—Snip restoration of cross section in Figure 2C to pre-Andean, post-Cretaceous state, in two segments with overlap
of central part, joining at X and X'. Black, Cretaceous–Paleogene sedimentary rocks; shaded, Paleozoic rocks and upper
crust; solid and dashed line, Moho. Foothills imbrication above Oriente foreland crust has remained unrestored for clarity. In
the computer-aided snip restoration, arbitrary graphic fragments are detached, translated, and rotated by simple shear. The
snip restoration then is cosmetically modified to restore graphic coherence.

the Eastern Cordillera is underlain by the upper crustal subduction and absence of volcanism. These conditions
shards of a Wernicke-type extensional structure are not evident in central Colombia.
(Wernicke, 1985). Possibly, the two separate crustal slabs In our own interpretation, foreland upthrusts may
overlying the eastern foreland are remnants of repeated have occurred in an early Neogene event of flat plate
events of low-angle extensional detachment. The poorly subduction. Parts of the Central Cordillera are deformed
constrained position of the Moho suggests a mantle in a style known from thin crust with a steep geotherm
upwarp, which is appropriate for an extensional center. (Isacks, 1988). An architecture of Mesozic and older
The restoration suggests that a restored length of 200 metasedimentary rocks involved in polyphase folding
km is affected by extensional strain (β) of about 2 and and soaked in post-kinematic granitoids is displayed on
ranging from 0 to 3. This lengthening would be consis- many Colombian geologic maps of the Central
tent with an extensional bulk strain on the order of Cordillera. Figure 10 shows a schematic cross section of
67–100 km. common style elements in flat plate subduction. The
most internal, or subductionward, part of this orogen
shows polyphase folds detached in the asthenosphere
below thin crust and lithosphere, described as type F
STRUCTURAL STYLE OF EASTERN basement structure (Hatcher and Hooper, 1992). Toward
CORDILLERA the external front, there are crustal thrust slabs described
as type C basement slabs, also detached in the litho-
In the Eastern Cordillera of central Colombia, sphere but with thicker, cooler upper parts. Farthest out
numerous outcrops of pre-Mesozoic tectonite are located is a field of foreland upthrusts detached in the lower or
within a field of very thick Mesozoic strata and are middle crust.
affected by thrust faults and several major mapped
strike-slip faults, not formally included in the present
analysis. The dip-slip components of Andean tectonics in
the Eastern Cordillera may be interpreted in two ways, COLOMBIAN ANDES: LITHOSPHERIC
and a choice is not possible from direct observation at CONFIGURATION
present. Either they reflect the preserved shoulders of
Mesozoic rift structures, or they suggest that the late Figure 11 is a composite lithospheric cross section of
Andean thrust transport was preceded by an event of the Andes in central Colombia. It shows a compilation
foreland upthrusts and mid-crustal detachment as and interpolation of the crustal and lithospheric data
described from Andean segments of flat plate subduc- available from the Andes in central Colombia. The cross
tion (Isacks and Barazangi, 1977; Isacks, 1988; Jordan et section shows the Peru-Chile subduction near its
al., 1983). There is some indirect support for the interpre- northern end (Meissnar et al., 1976). The complex slab of
tation as foreland upthrusts. Dengo and Covey (1993) the Panama-Carnegie plate mosaic has an estimated
invoke a late Andean architecture of foreland upthrusts elastic thickness of 28 km obtained from bathymetric
and a steep, abrupt fault limiting the Dahlstromian data (Lonsdale, 1978). A thermally defined base of litho-
foothills architecture. sphere is given at 70 km (Meissnar et al., 1976). The
As is understood at present (Isacks and Barazangi, Wadat-Benioff zone is in the “steep” category of the
1977; Isacks, 1988), upper crustal detachment and bimodal system (Isacks and Barazangi, 1977). The Andes
foreland upthrusts occur exclusively with flat plate may be underlain by a rising asthenospheric body if the
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution 643

TYPE F BASEMENT
ARCHITECTURE
TYPE C BASEMENT SLAB
FORELAND UPTHRUSTS

STRUCTURAL STYLE OF BASEMENT TECTONICS REFLECTS


GEOTHERM AND LITHOSPHERIC THICKNESS 100 KM

Figure 10—Sketch cross section of cratonic edge of an Andean orogen with flat plate subduction. Black and white,
supracrustal rocks; dark gray, crust; light gray, upper mantle; white dashed lines, brittle and plastic detachment surfaces.

WESTERN CORDILLERA CENTRAL CORDILLERA EASTERN CORDILLERA


TRENCH COAST CAUCA MAGDALENA CUSIANA
350 KM OFFSET

100 KM

BASE THERMAL LITHOSPH. ?

TOP OF CRUST
MOHO
100 KM
BASE OF FLEXURAL LITHOSPHERE

BASE OF THERMAL LITHOSPHERE

Figure 11—Schematic cross section of Andes and Peru-Chile subduction in central Colombia through latitudes 2°–4° N,
located as (3) in Figure 1 and based on Figure 2C (and on Case et al., 1973; Meissnar et al., 1976; Bermudez et al., 1985).
Black and white, supracrustal rocks; dark gray, crust; black line, Moho; light gray, upper mantle, with two definitions of the
base of lithosphere.

steepening took place during the Neogene. Its wall Acknowledgments Our interpretation has been reviewed
pressure along a horizontal density gradient is assumed and criticized by Mark Golborne, Scot Krueger, and Angela
(Le Pichon, 1983) to power the compression and root Thompson. This paper has received improvement through
formation in the Eastern Cordillera. However, this is not professional reviews and explanations by Jean Letouzey (Rueil-
supported by hard data. Also, the available density Malmaison) and Mark Cooper (Calgary). It has also received
gradient is only the minute and contested difference support from ARCO International and the German Research
between asthenosphere and lithosphere (of 0.05–0.1 Foundation (DFG). We thank them all.
g/cm3). More likely, therefore, sub-Andean thrusting is
powered by flow pressure in the rising and spreading
asthenosphere as it is pushed up ahead of deeper
material displacements.
REFERENCES CITED
Figure 11 shows undulations of the Moho at a half- Arango, J. L., T. Kassem, and H. Duque, 1976, Mapa
wavelength of 100 km or less. Therefore, an elastic load Geologico de Colombia: Bogota, INGEOMIN, scale
origin of this Moho curvature is unlikely. More likely, 1:1,500,000.
these roots are the result of crustal thrust faulting, as Bermudez, A., R. Acosta, and M. Garzon, 1985, Mapa Gravi-
shown in the cross section. metrico de Anomalias Simples de Bouguer: Bogota,
Figure 11 also shows the crust to contain a wide belt of INGEOMIN, scale 1:1,500,000.
Bally, A. W., P. L. Gordy, and G. A. Stewart, 1966, Structure,
accreted terranes in white, gray, and black, based on
seismic data and orogenic evolution of Southern Canadian
interpretations (Bourgeois et al., 1985; McCourt and Rocky Mountains: Bulletin of Canadian Petroleum
Aspden, 1985). Farther east, it shows the same succession Geology, v. 14, p. 337–381.
of flat slab styles as in Figure 10. Did the Eastern Bally, A. W., R. Catalano, and J. Oldow, 1985, Elementi di
Cordillera evolve from a field of foreland upthrusts? This Tettonica Regionale. Evoluzione dei Bacini sedimentari e
is possible and likely, but not proven. delle Catene Montuose: Bologna, Pitagorea Editrice, 276 p.
644 Roeder and Chamberlain

Bird, P., 1979, Continental delamination and the Colorado geodynamique depuis l’Eo-Cretace: 3rd cycle dissertation,
Plateau: Journal of Geophysical Research, v. 84, B13, University of Paris 6, Paris, 162 p.
p. 7561-7571. Heiskanen, W. A., and F. A. Vening Meinesz, 1958, The earth
Bourgeois, P., et al, 1985, Les ophiolites des Andes de and its gravity field: New York, McGraw-Hill, 470 p.
Colombie: evolution structurale et signification geody- Isacks, B., 1988, Uplift of the central Andean plateau and
namique (abs.): Memorias Congreso Latinoamericano de bending of the Bolivian orocline: Journal Geophysical
Geologia v. 6, p. 78. Research, v. 93, p. 3211–3231.
Brown, R. L., S. D. Carr, B. L. Johnson, V. J. Coleman, F. A. Isacks, B. L., and M. Barazangi, 1977, Geometry of Benioff
Cook, and J. L. Varsek, 1992, The Monashee décollement of zones: lateral segmentation and downwards bending of the
the southern Canadian Cordillera: a crustal-scale shear subducted lithosphere, in M. Talwani and W. C. Pitman III,
zone linking the Rocky Mountain foreland belt to lower eds., Island arcs, deep-sea trenches, and back-arc basins:
crust beneath accreted terranes, in K. R. McClay, ed., Thrust American Geophysical Union, Maurice Ewing Series, v. 1,
Tectonics: London, Chapman and Hall, p. 357–364. p. 99–114.
Butler, R. R., 1983, Andean-type foreland deformation: struc- James, D. E., 1971, Andean crustal and upper mantle structure:
tural development of the Neiva basin, Upper Magdalena Journal of Geophysical Research, v. 76, p. 3246–3271.
Valley, Colombia: Ph.D. dissertation, University of South Jordan, T. E., and R. W. Allmendinger, 1986, The Sierras
Carolina, Columbia, South Carolina, 272 p. Pampeanas of Argentina: a modern analogue of Rocky
Campbell, C. J., 1974, Colombian Andes, in A. M. Spencer, ed., Mountain foreland deformation: American Journal of
Mesozoic–Cenozoic orogenic belts: data for orogenic Science, v. 286, p. 737–764.
studies: Scottish Academic Press and Geological Society of Jordan, T. E., B. L. Isacks, R. W. Allmendinger, J. A. Brewer, V.
London, p. 705–724. A. Ramos, and C. J. Ando, 1983, Andan tectonics related to
Campbell, C. J., and H. Bürgl, 1965, Section through the geometry of subducted Nazca plate: GSA Bulletin v. 94,
Eastern Cordillera of Colombia, South America: GSA p. 341–461.
Bulletin, v. 76, p. 567–589. Karig, D. E., J. C. Caldwell, and E. M. Parmentier, 1976, Effects
Case, J. E., J. Barnes, G. Paris, H. Gonzales, and A. Vina, 1973, of accretion on the geometry of the descending lithosphere:
Trans-Andean geophysical profile, southern Colombia: Journal Geophysical Research, v. 81, p. 6281–6291.
GSA Bulletin, v. 84, p. 2895–2904. Le Pichon, X., 1983, Land-locked oceanic basins and conti-
Colletta, B., F. Hebrard, J. Letouzey, P. Werner, and J. L. nental Collision, in K. J. Hsu, ed., Mountain building
Rudkiewics, 1990, Tectonic style and crustal structure of the processes: London, Academic Press, p. 201–212.
Eastern Cordillera (Colombia) from a balanced cross Le Pichon, X., and J. C. Sibouet, 1981, Passive margins; a model
section, in J. Letouzey, ed., Petroleum and tectonics in of formation: Journal Geophysical Research, v. 86,
mobile belts: Paris, Editions Technip, p. 81–100. p. 3708–3720.
Dahlstrom, C. D. A., 1969, Balanced cross sections: Canadian Letouzey, J., 1990, Fault reactivation, inversion and fold-thrust
Journal of Earth Sciences, v.6, p. 743–757. belt, in J. Letouzey, ed., Petroleum and tectonics in mobile
Davis, D., J. Suppe, and F. A. Dahlen, 1983, Mechanics of fold belts: Paris, IFP Editions Technip, p. 101–128.
and thrust belts and accretionary wedges: Journal of Lonsdale, P., 1978, Ecuadorian subduction system: AAPG
Geophysical Research, v. 80, p. 1153–1172. Bulletin, v. 62, p. 2454–2477.
De Cizancourt, H., 1933, Tectonic structure of northern Andes Lyon-Caen, H., P. Molnar, and G. Suarez, 1985, Gravity
in Colombia and Venezuela: AAPG Bulletin, v. 17, anomalies and flexure of the Brazilian shield beneath the
p. 211–228. Bolivian Andes: Earth and Planetary Science Letters, v. 75,
Dengo, C.A., and M.C. Covey, 1993, Structure of the Eastern p. 81–92.
Cordillera of Colombia: implications for trap styles and Mattauer, M., 1986, Intracontinental subduction, crust-mantle
regional tectonics: AAPG Bulletin, v. 77, p. 1315–1337. décollement, and crustal-stacking wedge in the Himalayas
Dewey, J. F., M. R. Hempton, W. F. S. Kidd, F. Saroglu, andA. and other collision belts, in M. P. A. Coward and A. C. Ries,
M. C. Sengör, 1986, Shortening of continental lithosphere: eds., Collision tectonics: GSA Special Publication, v. 19, p.
the neotectonics of Eastern Anatolia—a young collision 37–50.
zone, in M. P. Coward and A. C. Ries, eds., Collision McCourt, W. J., and J. A. Aspden, 1985, Modelo tectonico de
tectonics: Geological Society of London Special Publication, placas para la evolucion Fanerozoica de Colombia central y
v. 19, p. 3–36. del sur: Memorias Congreso Latinoamericano de Geologia,
Estrada, J., 1982, Corte estructural a travez la Cordillera Bogota, v. 6, p. 1–35.
Oriental de Colombia, in R. Leigh, ed., Exploracion McKenzie, D., 1978, Some remarks on the development of
Petrolera de las Cuencas Subandinas, 3 vols., Bogota. sedimentary basins: Earth and Planetary Science Letters, v.
Giese, P., K. J. Reutter, V. Jacobshagen, and R. Nicolich, 1982, 40, p. 25–32.
Explosion-seismic crustal studies in the Alpine-Mediter- Meissnar, R. O., E. R. Flueh, F. Stibane, and E. Berg, 1976,
ranean region and their implications to tectonic processes, Dynamics of the active plate boundary in southwest
in H. Berckhemer and K. Hsu, eds., Alpine-Mediterranean Colombia according to recent geophysical measurements:
Geodynamics: Geodynamics Series, American Geophysical Tectonophysics, v. 35, p. 115–136.
Union, v. 7, p. 39–73. Miller, D. M., T. H. Nilsen, and W. L. Bilodeau, 1992, Late
Harding, T. P., and J. D. Lowell, 1979, Structural styles, their Cretaceous to early Eocene geologic evolution of the U.S.
plate tectonic habitats, and hydrocarbon traps in petroleum Cordillera, in B. C. Burchfiel, P. W. Lipman, and M. L.
provinces: AAPG Bulletin v. 63, p. 1016–1058. Zoback, eds., The Cordilleran orogen: conterminous U.S.:
Hatcher, R. D., Jr., and R. L. Hooper, 1992, Evolution of crys- GSA, The Geology of North America, v. G-3, p. 205–260.
talline thrust sheets in the internal parts of mountain Mitra, S., 1986, Duplex structures and imbricate thrust
chains, in , K. R. McClay, ed., Thrust tectonics: London, systems: geometry, structural position, and hydrocarbon
Chapman and Hall, p. 217–234. potential: AAPG Bulletin, v. 70, p. 1087–1112.
Hebrard, F., 1985, Les foot-hills de la Cordillere Orientale de Mitra, S., 1992, Balanced structural interpretations in fold and
Colombie entre les rios Casanare et Cusiana: Evolution thrust belts, in Shankar Mitra and George W. Fisher, eds.,
Eastern Cordillera of Colombia: Jurassic–Neogene Crustal Evolution 645

Structural geology of fold and thrust belts: Baltimore, Ulloa, M. C., and M. E. Rodriguez, 1978, Mapa geologico,
Johns Hopkins University Press, p. 53–77. Hoja de Chinquinquico: Bogota, INGEOMIN, 1:100,000.
Ocola, L. C., L. T. Aldrich, J. F. Gettrust, R. P. Meyer, and J. E. Valderrama, R., 1982, Desarollo de facies en la cuenca de los
Ramirez, 1975, Project Narino I: crustal structure under Llanos Orientales Colombianos, in R. Leigh, ed., Explo-
southern Colombian–northern Ecuador Andes from racion Petrolera de las Cuencas Subandinas, 3 vols.,
seismic refraction data: Bulletin of the Seismological Bogota.
Society of America, v. 65, p. 1681–1695. Vann, I. R., R. H. Graham, and A. B. Hayward, 1986, The
Oxburgh, E. R., and D. L. Turcotte, 1974, Geodynamic inter- structure of mountain fronts: Journal of Structural
pretation of Tauern metamorphism, eastern Alps: Geology, v. 8, p. 215–228.
Schweizer Mineralogische und petrographische Watts, A. B., and W. B. F. Ryan, 1976, Flexure of the lithos-
Mitteilungen, v. 52, p. 479–495. phere and continental margin basins: Tectonophysics, v.
Ramsay, J. G., and M. I. Huber, 1987, The techniques of 36, p. 25–44.
modern structural geology, vol. 2, Folds and fractures: Wernicke, B. P., 1985, Uniform-sense normal simple shear of
Orlando, Florida, Academic Press, 700 p. the continental lithosphere: Canadian Journal of Earth
Renzoni, G., H. Rosas, and F. Etayo-Serva, 1983, Mapa Science, v. 22, p. 108–125.
geologico, Hoja de Tunja: Bogota, INGEOMIN, scale Williams, G. D., C. M. Powell, and M. A. Cooper, 1989,
1:100,000. Geometry and kinematics of inversion tectonics, in M. A.
Roeder, D., 1989, South-Alpine thrusting and trans-Alpine Cooper and G. D. Williams, eds., Inversion tectonics:
convergence, in M. P. Coward, D. Dietrich, and R. G. Park, Geological Society of London Special Publication 44,
eds., Alpine tectonics: Geological Society of London p. 3–15.
Special Publication 45, p. 211–242. Ziegler, P. A., 1990, Geological atlas of western and central
Roeder, D., 1991, Compressional tectonics and the balancing Europe, second edition: Shell Internationale Petroleum
of crustal cross sections, in P. Giese, D. Roeder, and R. Maatschappij B.V. and Geological Society of London,
Nicolich, eds., Joint Interpretation of geophysical and 216 p.
geological data applied to lithospheric studies: Dordrecht,
The Netherlands, Kluwer Academic Publishers, Nato AST
Series 338, p. 127–163.
Royden, L., and G. D. Karner, 1984, Flexure of the continental
lithosphere beneath Apennine and Carpathian foredeep
basins: evidence for an insufficient topographic load:
AAPG Bulletin, v. 68, p. 704–712.
Shirley, K., 1992, Cusiana a world-class discovery: Colombia
Authors’ Mailing Addresses
find wows explorers: AAPG Explorer, August, p. 1, 16–19. D. Roeder
Suppe, J., 1983, Geometry and kinematics of fault-bend Institut für Lithosphärenforschung
folding: American Journal of Science, v. 283, p. 684–721. Justus Liebig–Universität, Senckenbergstr. 3
Sobornov, K. O., 1991, The formation of fold-nappe structure D-35390 Giessen
of the Dagestan Wedge: Geotektonika, v. 3, p. 34–46. Germany
Trümpy, R., 1980, Geology of Switzerland, part A: an outline
of the geology of Switzerland; part B: geological excur-
sions: Basel, Schweizerische Geologische Kommission, R. L. Chamberlain
104 p. Blue Eagle Exploration Inc.
Turcotte, D. L., and G. Schubert, 1982, Geodynamics; applica- 625 Green Road
tion of continuum physics to geological problems: New Salisbury, North Carolina 28147
York, John Wiley, 450 p. U.S.A.
Geodynamic Evolution of the Eastern Andes,
Colombia—An Alternative Hypothesis

Peter B. Jones
International Tectonic Consultants Ltd.
Calgary, Alberta, Canada

Abstract

R ecent hypotheses of the Miocene–Pliocene structural evolution of the Eastern Cordillera of the Andes in
Colombia have proposed a divergent fan model combining a system of southeast-verging, basement-
rooted thrust faults beneath the eastern edge of the Cordillera with a system of northwest-verging thrusts
rooted in the basement beneath its western edge. However, structural relationships within the western thrust
system indicate that its faults are superficial posttectonic gravity-driven features. This leads to an alternative
hypothesis for the evolution of the Colombian Andes, which has significant structural, stratigraphic, and
temporal implications for hydrocarbon exploration.
In the new model, the Eastern Cordillera and Perija Andes are interpreted as elements of a single
southeast-verging thrust sheet with minor imbrications. Initially, this thrust sheet moved east-southeastward
as a blind thrust sheet with triangle zone (buried thrust front) geometry, creating the continuous uplifted area
of the Eastern Cordillera, Santandar massif, and Perija Andes. The overthrust sheet was subsequently trans-
ported “piggy-back” farther southeast to its present location by emplacement of the underlying Merida
Andes thrust system, for a total shortening of approximately 185 km. According to the new model, the
incipient leading edge of this overthrust sheet was offset by an oblique lateral ramp. After deformation, the
resultant lateral step in the footwall was accommodated by normal fault movement in the overthrust sheet,
forming the Bucaramanga fault, with an apparent strike-slip offset of 210 km.

Resumen

H ipótesis recientes relativas a la evolución estructural experimentada durante el período Mio-Pliocénico


por la Cordillera de los Andes Orientales en Colombia, han sugerido un modelo de abanico divergente
que combina un sistema suroriental de fallas de caballamiento que originaron en el fundamento, ubicadas bajo
el límite oriental de la Cordillera, con otro sistema de fallas proyectado hacia el noroeste, situadas en funda-
mentos bajo el límite occidental. Sin embargo, las relaciones estructurales encontradas en el interior del sistema
de fallas occidentales indican que las mismas poseen características superficiales, originadas por desplaza-
mientos gravitacionales de la corteza terrestre. Esta situación permite formular una hypótesis alternativa para
explicar la evolución de los Andes Colombianos, lo cual pose importantes implicaciones estructurales, estrati-
graficas y temporales aplicables a la exploración de yacimientos de hidrocarburos.
En el nuevo modelo, la Cordillera Orientale y los Andes de Perija son considerados como elementos de una
falla única, situada en la dirección sureste, y ubicada en depósitos sedimentarios que exhiben un traslapo
mínimo. Inicialmente, la falla se desplazó en la direccion esta - sureste, como una falla escondida (de frente
subterráneo), creando la zona de elevacíones continuas que contiene la Cordillera Oriental, el Macizo de
Santandar, y los Andes de Perija. El cuerpo de roca que estaba ubicado sobre la falla fueron subsecuentemente
transportados más hacia el sureste, a su actual ubicación, por la intrusión del sistema de fallas subterráneas de
los Andes de Mérida, produciéndose una modificación de cerca de 185 kilómetros. De acuerdo con el nuevo
modelo, el borde inicial de esta capa, ubicada sobre la falla, fué afectado por una distorsión lateral oblicua. Con
posterioridada a la formación de esta distorsión, la discontinuidad producida en la pared rocosa fué compen-
sada por el movimento natural experimentado por la corteza situada sobre la falla, formandóse así la fractura
de Bucaramanga, con un desplazamiento aparente de 210 kilómetros.

Jones, P. B., 1995, Geodynamic evolution of the eastern Andes, Colombia—an alternative 647
hypothesis, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 647–658.
648 Jones

Figure 1—Location map showing


tectonic elements of the
Colombian and Venezuelan
Andes described in the text.
Location of the cross section in
Figure 11 is shown by the double
line. The dashed box outlines the
area of Figure 5.

INTRODUCTION (Jones, 1993). It describes three interrelated critical factors


that affect both the amount of movement and the
Hypotheses concerning the manner and estimates of manner in which it occurred, with major implications for
the extent of Miocene–Pliocene deformation of the the petroleum geology of the region. The first is the
Eastern Cordillera of Colombia, (Figure 1) have varied evidence that the west-verging thrusts in the Magdalena
considerably. Irving (1975) attributed the formation of basin are superficial gravity-driven structures, conse-
the Cordillera to vertical uplift, and his cross sections quent upon but otherwise irrelevant to the basic struc-
show all faults within or flanking the Cordillera dipping tural evolution of the Andes. The second is the existence,
at 60° or greater. Cross sections by Butler and Schamel on varying scales, of buried thrust fronts or triangle zones
(1988) also show steeply dipping faults along the east (Gordy et al., 1977; Jones, 1982, 1987; Teal, 1983; Morley
and west margins of the Cordillera, implying mainly 1986; Vann et al., 1986), where thrust slip is not trans-
vertical uplift of 5–20 km. Laubscher (1987), who invoked mitted toward the foreland but rather absorbed along
low-angle thrusting as the mechanism of uplift, calcu- one or more detachment zones, uplifting the overlying
lated, on the basis of the uplifted volume of the Eastern sequence, somewhat in the manner of a horizontal
Cordillera, that southeastward horizontal displacement igneous intrusion. This process has been demonstrated
of about 150 km had taken place along low-angle thrusts. along the foreland margins of almost all Phanerozoic
Although recent interpretations of the structure have thrust and fold belts. The third factor is an alternative
involved low-angle thrusts (Lausbscher, 1987; Colletta et interpretation of the nature of the Bucaramanga fault.
al., 1990; Dengo and Covey, 1993), these authors have
postulated the presence of basement-rooted southeast-
verging thrust faults beneath the eastern edge of the GRAVITY-DRIVEN STRUCTURES
Cordillera, combined with northwest-verging thrust
faults, also rooted in the basement in a fanlike configura- Geometry of a Gravity-Driven Slip Sheet
tion. They correlate the northwest-vergent thrust system
with northwest-vergent thrust faults that crop out in the Gravity-driven structures range in size and area from
eastern part of the Magdalena Valley, which parallels the outcrop-scale structures to entire deformed belts. They
western slope of the Eastern Cordillera. Estimates of the occur where there is sufficient topographic relief to
total minimum horizontal movement of this bilateral provide a driving force and the rocks are incompetent
combination range from 100 km (Colletta et al., 1990) to enough to facilitate slippage. Dip slopes of layered sedi-
150 km (Dengo and Covey, 1993). mentary rocks provide ideal conditions for down-dip
This paper is expanded from an oral presentation sliding. Most of the movement is within less competent
Geodynamic Evolution of the Eastern Andes, Colombia—An Alternative Hypothesis 649

limb corresponds to the head of the slide, striking parallel


to the edge of the uplift. The other limb is oblique to the
uplift and includes the overthrust sector and thrust front.
Because of the plunge, the thrust in the oblique limb is
emergent near the fold axis (Figure 2), becoming buried
away from the axis and down the plunge (Figure 3).

Gravity-Driven Slip Sheets in the


Figure 2—Diagram of gravity-collapse structures. Magdalena Valley
(Redrawn from Harrison and Falcon, 1936.) Examples of
these structures on different scales have been described in A belt of north-plunging synclines extends along a
the Colombian Andes by Julivert (1970). The slip sheet in 450-km stretch along the eastern side of the Magdalena
this diagram has an emergent thrust front in contrast to the
Valley from the Bucaramanga fault to the constriction of
slip sheet in Figure 3, which has a buried (triangle zone)
thrust front. the valley southwest of Bogota (Figure 1), which marks a
boundary between subbasins. A tectonic map of a sector
of the northern part of the belt (Figure 5) illustrates the
stratigraphic units, parallel or subparallel to the bedding. characteristic pattern. In previous interpretations of the
Gravity-driven structures include cascade folds, flaps, structure of the Eastern Cordillera (Laubscher, 1987;
and slip sheets (Figure 2). Slip sheets may form synclines Colletta et al., 1990; Schamel, 1991; Dengo and Covey,
striking parallel or subparallel to dip slopes flanking 1993), the west-verging thrusts underlying the western
major uplifts (Julivert, 1970; Jones, 1988, 1990). limbs of these synclines were interpreted as basement-
The toe (lower end or leading edge) of a slip sheet has rooted tectonic features (Figures 5a, 6a). Geometry
a thrust front that may be exposed (Figure 2) or buried at negates this interpretation and suggests their reinterpre-
depth, with a frontal triangle zone (Figure 3) involving tation in the manner of Figures 5b and 6b, which incor-
one or more blind thrust faults. From the thrust front, the porates the slip sheet geometry of Figures 2, 3, and 4b,
basal thrust follows the bedding beneath the synclinal and does not require any direct basement involvement.
axis and up the slope to the head area (Figure 3). Grabens Section missing along the eastern limbs of the synclines
and other associations of normal faults and omission of indicates low-angle normal faulting at the head areas of
section are normal features of the head of a gravity slide gravity-driven slip sheets.
(Graham, 1981, Jones, 1990), distinguishing a slip sheet The Las Salinas thrust fault (Figure 5) underlies Upper
from a folded thrust of tectonic origin. Anticlinal struc- Cretaceous Umir shale in the western flank of the most
tures in the eastern part of the Magdalena Valley may northerly of these synclines, the Nuevo Mondo syncline
overlie toe structures similar to those in Figure 3. (Figure 5b). In the eastern flank of the syncline, the
Section missing at the head of a slip sheet may be due Lebrija fault also follows the Umir shale (Ward et
to a small fault perpendicular to bedding (Figure 4a) or al.,1977). It can be correlated with the Las Salinas fault
to a major fault subparallel to bedding (Figure 4b). If the along the bedding of the Umir shale (Figure 5b). Several
fault consistently follows one stratigraphic unit, the inter- workers (e.g., Schamel, 1991) have traced this major east-
pretation in Figure 4b is far more reasonable and dipping bedding-parallel thrust fault southward into a
provides a link to toe structures. The fault in Figure 4a is minor bedding-perpendicular fault that extends along
relatively trivial. the eastern edge of the Magdalena Valley and follows the
A combination of plunge and topography may result trace of the Upper Cretaceous shale unit along the
in a synclinal slip sheet on the flank of an uplift having common eastern flank of several synclines (Figure 6a).
the V-shaped outcrop pattern of a plunging syncline. One However, the geometric constraints previously outlined

Figure 3—Cross-section of a
slip sheet with a buried (triangle
zone) thrust front, northern Tien
Shan mountains, Xinjiang,
China, about 250 km west of
Urumchi. Multiple thrusts in the
toe of the slip sheet pass up
into normal faults at its head.
This structure lies at the
northern (trailing) edge of the
south-vergent Tien Shan thrust
system. (From Jones, 1990.)
Similar structures and pro-
cesses have been described by
Graham (1981) and Gramond
(1993) at leading edges of uplifts
in the French Alps.
650 Jones

The pattern of north-plunging synclines is remarkably


consistent and repetitive (Figure 5). A northeast-
southwest cross section along the axes of the synclines
(Figure 7a) resembles a cross section across a typical
thrust belt and indicates a sequence of emplacement of
slip sheets from higher (north) to lower (south). This
sequence may be related to southward propagation of
the Merida Andes beneath the Eastern Cordillera during
the latest stage of formation of the Eastern Cordillera,
uplifting the northern area first, close to the Bucara-
manga fault. Regardless of where it was located, an east-
west cross section through the synclinal belt would look
like Figure 7b, intersecting two or more plunging slip
sheets at different levels, the deeper ones having buried
thrust fronts with triangle zone geometry and the upper
ones with emergent thrust fronts.
Small-scale gravity-driven structures in the Eastern
Cordillera have been described by Julivert (1970). They
include flap structures, cascade folds, and synclinal slip
sheets (Figure 2) similar to but much smaller than the
Nuevo Mondo and other regional synclines. The struc-
tural relief along the eastern side of the Magdalena basin
is about 10–12 km (Irving, 1975), and the present physio-
graphic relief is about 3 km. Although there is consider-
able variation, this structural relief created dip slopes
along the eastern side of the Magdalena Valley with
westward dips of 10°–40°, angles that are more than
adequate for large-scale gravity sliding westward,
possibly generating still unrecognized flap structures
and cascade folds in addition to slip sheets.
Although the Cesar Valley (Figure 1) has not been
examined in the same detail, it is probable that rootless
west-verging thrust structures observed on seismic
profiles across the eastern part of the valley are also the
result of westward gravity sliding down the trailing edge
Figure 4—Alternative interpretations of the same outcrop of the Perija overthrust sheet.
data. (a) Small extensional fault subperpendicular to
bedding of a dip slope. Slip of the fault is approximately
equal to the stratigraphic offset across it. Compared to the
amount of fault movement involved in producing the major BLIND THRUSTING IN COLOMBIA-
structure on whose flank the fault occurs, such a fault is
insignificant. (b) The extensional fault subparallel to VENEZUELA SECTOR OF EASTERN
bedding has the same stratigraphic offset as in part (a) and ANDES
may be misinterpreted in that way. However, slip subpar-
allel to bedding can accommodate a major thrusting The second of the three interrelated factors bearing on
downslope as in Figure 3. the geodynamics of the Eastern Cordillera is the presence
of buried thrust fronts (Morley, 1986), passive roof duplexes
(Banks and Warburton, 1986), or triangle zones (Gordy et
do not support this style of interpretation or correlation. al., 1977), alternative terms for the same structure. They
At its northern end, the Lebrija fault swings northwest occur along the foreland margins of most of the world’s
parallel to the Bucaramanga fault, as does the axis of the mountain belts. In a buried thrust front, blind thrusting is
Nuevo Mondo syncline, and both pass beneath Quater- accommodated along an upper detachment or passive
nary valley fill. roof thrust (Jones, 1982; Banks and Warburton, 1986),
The reinterpretation shown in Figure 6b, involving a sometimes referred to as a back-thrust. The section above
fault parallel and therefore common to the opposing the detachment surface is differentially uplifted and
flanks of a syncline, can be applied to most of the other folded in harmony with the structures in the underlying
synclinal structures in the eastern part of the Magdalena thrust duplex. If syntectonic erosion through the upper
Valley. Clearly, the gravity sliding shown in Figure 6b detachment prevents continued blind thrusting,
could occur within the hanging wall syncline of Figure continued compression may be accommodated by thrust
6a, but while the gravity-sliding model could account for faults arising from the upper detachment either as
the structures in Figure 5 in their entirety, the tectonic foreland-vergent emergent thrusts or as hinterland-
thrusting of Figure 6a alone cannot. verging back-thrusts (Banks and Warburton, 1986; Vann
Geodynamic Evolution of the Eastern Andes, Colombia—An Alternative Hypothesis 651

Figure 5—(a) Regional structure, Middle Magdalena Valley, Colombia, showing a suite of north-plunging synclines. (From
Taborda, 1965; Schamel, 1991.) (b) Same area with gravity slide interpretation of synclines along the eastern edge of the
basin. Note the correlation of the Las Salinas thrust with the Lebrija normal fault. High ground is indicated by horizontal
pattern. Location of map area is shown in Figure 1.

et al., 1986). It is not known whether such faults reached southeastward for over 500 km and dies out in the
the surface or flattened in a higher detachment level at Eastern Cordillera. Topographic and geologic maps
the time of their emplacement that has subsequently show an obvious displacement of the Cesar Valley
been removed by erosion. relative to the Magdalena Valley across the Bucaramanga
Buried thrust fronts have been recognized along long fault. This has led most authors (e.g., Dengo and Covey,
stretches of the eastern margin of the Andean chain from 1993) to attribute this offset to left-lateral strike-slip fault
Tierra del Fuego to Venezuela, including the Perija movement of about 210 km. A major problem inherent in
Andes in Venezuela and Colombia (Audemard, 1993), this interpretation is that geologic maps (Ingeominas,
Cordillera de la Costa, Venezuela (Bell and Jones, 1993), 1988; Colletta et al., 1990; Schamel, 1991) show that the
and Merida Andes (Urbina, 1993). A foreland-dipping offset appears to die out completely in a southeastward
monoclinal mountain front characteristic of buried thrust direction within the Eastern Cordillera over a distance of
fronts extends along most of the eastern margin of the only 75 km. It is not possible for that amount of strike-
Eastern Cordillera. In some sectors of the foothills, slip movement to die out over such a short distance.
thrusts such as the Guicarama, Yopal, and Cusiana Even if some of that movement was accommodated by
thrusts in the Llanos foothills may represent reactivation west-verging thrusts within the Cordillera, as suggested
of the originally buried thrust front, rooted in the Upper by Dengo and Covey (1993), most of the movement is
Cretaceous shale section that provided the upper detach- still unaccounted for. However, interpretation of the
ment zone for the original blind thrusting. Bucaramanga fault as a normal fault (Figure 8) accom-
modating the vertical displacement of an underlying
lateral ramp eliminates that problem. Movement on the
NATURE OF THE BUCARAMANGA fault would be mainly vertical, 10 km or less, an amount
FAULT that can die out along strike over a short distance. The
apparent offset is a function of the shape of the original
The third element in this tectonic reappraisal involves Perija–Eastern Cordillera overthrust sheet, and the
the Bucaramanga fault, a regional linear near-vertical amount of apparent offset corresponds to the length of
fault that extends from the Caribbean coast of Colombia the original lateral ramp.
652 Jones

Figure 10a shows the configuration of the Perija–


Eastern Cordillera thrust sheet prior to movement. In
Figure 10b, movement is shown to be southeastward,
above the frontal ramps and oblique to the lateral ramp.
This resulted in uplift of the Santandar massif along with
the Eastern Cordillera and Perija Andes. Although both
of these ranges are thrust faulted internally, these faults
are relatively small compared to the 50–60 km of
movement required to create the uplifts as a whole. The
second stage of movement (Figure 10c) arose from
emplacement of the Merida Andes beneath the
combined Perija–Eastern Cordillera thrust sheet, carrying
it southwestward a further 90 km. It is possible that as it
propagated southward, the sole thrust of the Merida
Andes passed upsection into a bedding plane in the
same Upper Cretaceous shale interval that formed the
glide horizon beneath the Eastern Cordillera. Extrapo-
lated southward, the footwall ramp of the Merida Andes
could have been joined with the footwall ramp of the
Eastern Cordillera sole thrust subjacent to the eastern
side of the Magdalena basin.
Figure 11 is a series of retrodeformable cross sections
showing the tectonic development of the Eastern
Cordillera. They are based on the cross section by
Figure 6—Subsurface relationships implicit in the different
Colletta et al. (1990). They were simplified to show how
interpretations of Figure 5. (a) Cross section through a
typical syncline according to Schamel (1991). This subsur- the same surface structures could be produced with east-
face interpretation, which involves a northwest-vergent dipping thrusts only. Figure 11b illustrates the effect of
thrust rooted in a mid-crustal detachment is not compatible blind thrusting in the initial stages of development of the
with the consistent presence of faults in the eastern flank Eastern Cordillera–Perija thrust sheet. The sole thrust
of synclines in Figure 5a. It is also clear that if there were a follows a mid-crustal detachment (Figure 6), ramping
southeast-dipping fault cropping out in the southeastern upward to a detachment in Upper Cretaceous shales.
limb of the syncline, it could not merge with the thrust in This section has different names in different localities,
the northwestern flank. Also, according to this model, the including Umir (northern Magdalena Valley), Villeta
sections southeast and northwest of the fault-bend fold (southern Magdalena Valley), and Monserrate (Llanos
(i.e., the Eastern Cordillera) would have to lie at similar
foothills). Southeastward movement of the blind thrust
structural levels. In the Magdalena Valley, however, Creta-
ceous and older rocks in the Cordillera are more than 10 sheet between these two detachment surfaces (Figure
km higher than their equivalents beneath the Magdalena 11b) continued until erosion removed the cover and
Valley. (b) This interpretation accounts for the consistent exposed the upper detachment, breaching the under-
presence of a fault in the southeastern limb of each lying thrust duplex.
syncline that can be correlated directly with the thrust in Blind thrust duplexes or triangle zones are typically
each northwestern limb. The underlying southeast-vergent overpressured (Suppe, 1985), and removal of that over-
thrust rooted in the mid-crustal detachment can be related pressure inhibits the blind thrusting process. With the
directly to thrust emplacement uplifting the Eastern upper detachment inactive, continued compression was
Cordillera. taken up by thrusts in the Llanos basin, which may have
been related to movement of the Merida Andes. The
aggregate slip of these thrusts was too small to have
A NEW MODEL FOR THE STRUCTURAL moved the Eastern Cordillera by more than a few tens of
kilometers. However, if the bulk of southeastward
EVOLUTION OF THE EASTERN displacement occurred through blind thrusting, then the
CORDILLERA Llanos basin thrusts (such as the Yopal and others)
added their slip to the preexisting and much greater
Elimination of west-verging thrusts in the Middle displacement of the initial blind thrust sheet. Rooted in
Magdalena Valley from the tectonic model permits a Upper Cretaceous shale, it is possible that these thrusts
simpler model of the tectonic evolution of the eastern pass westward and northward into the sole fault of the
Andes of Colombia, involving southeastward thrusting Merida Andes.
instead of the earlier divergent fan models. This model Syntectonic erosion of the uplifted shallow section
treats the Eastern Cordillera and Perija Cordillera as means that it was in a tensional environment above the
elements of a single continuous thrust sheet whose original footwall thrust ramp, a condition that favored
leading edge includes two frontal ramps connected by a westward-vergent gravity sliding above the ramp along
lateral ramp (Figures 9, 10). The geodynamic develop- the eastern edge of the Magdalena Valley. Figures 5 and
ment of the orogen is illustrated in Figure 10. 7 show how each slip sheet plunges beneath the one to
Geodynamic Evolution of the Eastern Andes, Colombia—An Alternative Hypothesis 653

Figure 7—(a) Diagrammatic


strike cross section along the
synclinal axes of Figure 5,
showing the repetitive pattern of
slip sheet emplacement and rela-
tionships between successive
slip sheets within the southward-
deepening Magdalena basin.
Although the cross section
resembles that across a typical
thrust belt, movement is mainly
out of the plane of the diagram
(westward). (b) Typical cross
section showing the association
of buried and emergent thrust
fronts in the Magdalena basin.
Each emergent thrust front
passes down-plunge northward
into a buried thrust front in the
west limb of each syncline,
creating a north-plunging
anticline in the overlying section.
Location of the longitudinal
cross section is indicated by the
vertical dashed line.

the north of it, suggesting that they are progressively


younger to the south. The initial movement may have
been triggered by the last stages of movement of the
Bucaramanga fault, which marks the northern limit of
the slip sheets. Another factor causing the north-south
sequence of emplacement may be the progressive
southward propagation of the Merida Andes thrust
system, uplifting the Eastern Cordillera and transporting
it farther to the southeast. It may not be coincidental that
the present summit of the Eastern Cordillera–Perija
mountain chain overlies the southward projection of the
Merida Andes.
Following the model just described, it is possible to
determine the amount of horizontal displacement of the
eastern edge of the Cordilleran thrust sheet by Figure 8—Diagram showing how horizontal offset can be
produced by either normal or strike-slip fault movement.
measuring the distance in the direction of tectonic However, strike-slip movement must extend beyond the
transport (northwest-southeast) from the top of the ramp region of offset or be accommodated in some way. Normal
on the eastern side of the Magdalena Valley to the fault movement is relatively small and can vary over short
present buried thrust cutoff of the Paleozoic–Lower distances
Cretaceous section at the front of the Eastern Cordillera,
a distance of about 150 km. This does not take into and associated structures in the southern Canadian
account the thrusting within the Eastern Cordillera itself, Cordillera but on a larger scale (Figure 12). The Lewis
which may involve an additional amount of shortening thrust is subhorizontal to gently folded over a wide area
along a number of thrusts, including the Chiveta, Cormi- and has thrust Proterozoic clastics and carbonates more
choqua, and Santa Barbara faults (Colletta et al., 1990) or than 60 km eastward over Mesozoic clastics (Gordy et al.,
the Boyaca, Soapaga, and Chameza faults (Dengo and 1977). The overthrust sheet was subsequently trans-
Covey, 1993). Thus, this model suggests a total hori- ported piggy-back another 50 km southeastward by
zontal displacement of about 185 km, an amount that is thrusting in duplex structures beneath it and within the
within the range obtained by previous workers, although foothills belt ahead of it.
by a very different and simpler tectonic process. The Lewis thrust sheet includes a major lateral ramp
that offsets the Clark and Flathead ranges, analogs of the
Comparison with the Lewis Thrust Sheet of Eastern Cordillera and Perija Andes. Beneath it is the
the Canadian Cordillera south-plunging end of the Livingstone thrust sheet, a
small-scale analog of the Merida Andes. West of the
The structure of the Colombia-Venezuela sector of the Lewis thrust is the Flathead normal fault system, which
eastern Andes is similar to that of the Lewis thrust sheet parallels the Lewis thrust front. Over most of its course,
654 Jones

the subjacent foothills structures, the Cordillera Orientale


thrust sheet moved a total of 185 km, but the narrow sub-
Andean foothills thrust belt ahead of it suggests that the
contribution from foothills thrusting was relatively small.

IMPLICATIONS FOR HYDROCARBON


EXPLORATION
Mesozoic Facies Distribution
This model of thrusting of the Eastern Cordillera has
Figure 9—Geometry of a lateral ramp within an overthrust consequences for the predeformation distribution of
sheet. Movement oblique to the strike of the lateral ramp sedimentary facies in Lower Cretaceous and older sedi-
produces transpressional structures at the leading edge of mentary rocks. The thick Lower Cretaceous section in the
the oblique sector of the thrust sheet. Following
Magdalena Valley is absent from the Llanos foothills,
movement, tensional stress above the footwall lateral ramp
may cause normal faulting in the hanging wall. (Modified where Upper Cretaceous strata rest on Paleozoic or older
from McClay, 1992.) rocks. Previous geodynamic models of the Eastern
Cordillera, involving smaller amounts of movement on
the Eastern Cordillera sole thrust, allow a limited range
the Flathead fault dips westward, generally subparallel of possible facies models and locations of the eastern
to bedding, except where it follows the underlying lateral edge of the sub–Upper Cretaceous section. The new
ramp of the Lewis thrust and dips steeply southward. In model is compatible with an eastern limit anywhere
that sector, it resembles the Bucaramanga fault on a between the Magdalena Valley and the western edge of
smaller scale, with apparent strike-slip offset between the the Llanos basin, beneath the Eastern Cordilleran over-
Fernie and Flathead basins, small-scale analogs of the thrust sheet. One of the possibilities is shown in Figure
Cesar and Middle Magdalena basins, respectively. 11, in which a faulted eastern boundary to the Lower
Unlike the Lewis thrust sheet, which moved about 60 km Cretaceous basin is shown. It is also possible that the
and was transported piggy-back an additional 50 km by ramp for thrust emplacement of the Eastern Cordillera

Figure 10—Sketch maps to show the tectonic evolution of the Colombian Andes. The marker (+) in the lower right corner of
each panel indicates a fixed reference point. (a) Before movement, showing the Perija and Eastern Cordilleran sectors
connected by a lateral ramp. Arrows indicate direction of movement. (b) East-southeastward movement of the combined
thrust sheet oblique to the lateral ramp, producing a transpressional structural environment northeast of the Santandar
Massif and a tensional environment above the footwall lateral ramp. (c) Southeastward transport of the combined sheet by
the underlying Merida Andes thrust system to its present position.
Geodynamic Evolution of the Eastern Andes, Colombia—An Alternative Hypothesis 655

Figure 11—Kinematic model of the evolution of the Cordillera Orientale. These computer-synthesized balanced cross
sections were created on equal horizontal and vertical scales but have been compressed 5:1 horizontally for space reasons.
(a) Before movement, showing location of major faults. The normal fault in the Lower Cretaceous is speculative, inserted to
suggest a range of possibilities of the configuration of the Lower Cretaceous basin. (b) Blind thrusting stage, with buried
thrust front (triangle zone geometry) at the leading edge of the advancing blind thrust sheet. Erosion through the uplifted
roof section breached the upper detachment zone within the Upper Cretaceous shale. Loss of overpressure forced transfer
of displacement to the foothills thrusts. (c) Present configuration. Nature of Llanos foothills front is uncertain. Slip sheet in
the Magdalena Valley has created the Cimitarra syncline and San Fernando anticline to the west of it.
656 Jones

Figure 12—Comparison of (a) the Colombia Andes and (b) the southern Canadian Cordillera. In the Canadian example, the
Flathead and Clark ranges are analogs of the Perija and Eastern Cordilleras, respectively. The Livingstone Range is
analogous to the Merida Andes. Gravity sliding occurs in both regions. Dashed lines indicate structural trends. (From Gordy
et al., 1977; Ingeominas, 1988.)

thrust sheet was formed by inversion of the earlier fault. through some parts of the synclinal slip sheets should
It is also possible that the Lower Cretaceous section is encounter low-angle normal faults downthrown to the
truncated gradually eastward by the Upper Cretaceous west, with consequent omission of section. This appears
unconformity. to have happened in the Colorado field (Figure 13),
where the Tertiary Chorro Formation directly overlies a
Structural Traps variety of older rocks. In the original interpretation of
this structure (Taborda, 1965), the contact between the
Structural traps in the Magdalena basins occur in Chorro and the underlying Cretaceous section was inter-
strike-slip fault related structures as well as in thrusted preted as an angular unconformity. However, this does
Cretaceous and Tertiary strata. The thrust-related struc- not explain why the upper section is tightly folded while
tures can be divided into two types according to the the lower section is not, a reversal of the usual relation-
proposed model: those overlying east-verging thrust ship across an angular unconformity. If the folding were
faults, such as the Dina and San Francisco fields of the restored to a horizontal predeformation state, the under-
Upper Magdalena Valley, and those associated with lying section would be restored from its present gently
west-verging gravity slides. Although the latter should tilted state to a tightly folded predeformation condition,
have a lower hydrocarbon potential because of their late which is difficult to accept. If that Cretaceous-Tertiary
formation and limited catchment area, several major contact is a low-angle normal fault, the relationships are
accumulations occur in the western limbs of synclines more easily reconciled.
(Schamel, 1991), including the Provincia field in the
Nuevo Mondo syncline.
CONCLUSIONS
Low-Angle Normal Faults in the
Magdalena Valley Evidence of a posttectonic gravitational origin for
northwest-vergent thrust faults in the Magdalena Valley
If westward gravity sliding is responsible for some of questions the validity of existing bilateral thrust models
the structures in the Magdalena basin, wells drilled of the Eastern Cordillera and leads to the proposal of a
Geodynamic Evolution of the Eastern Andes, Colombia—An Alternative Hypothesis 657

Figure 14—Block diagram illustrating the results of south-


eastward thrusting of the combined Perija–Eastern
Cordillera thrust sheet, gravity sliding in the Magdalena
Figure 13—East-west cross section through the Colorado Valley, and vertical movement of the Bucaramanga fault.
field, Magdalena basin. In this interpretation, a low-angle
normal fault downthrown to the west replaces the angular
unconformity originally interpreted to lie at the base of the The footwall ramp of the Eastern Cordillera–Perija
Chorro (Taborda, 1965). The reinterpretation explains why thrust sheet now underlies the eastern flanks of both the
the Tertiary Chorro-Real section above the break is folded,
while the underlying Jurassic–Cretaceous Giron through
Magdalena and Cesar basins. The flat underlies the
the Rosa Blanca section is not, a reversal of normal uncon- Eastern Cordillera, Santandar massif, and Sierra de
formable relationships. Location of cross section is given Perija, and the tip lines of the thrust system, typically
by CF in Figure 5b. (Modified from Taborda, 1965.) buried, lie east of the Eastern Cordillera, Perija, and
Merida Andes.
The effects of this new model on the hydrocarbon
new model for its geodynamic evolution. According to potential include possible changes in relationships
the new model, the Eastern Cordillera and Perija among source and reservoir facies, migration paths,
Cordillera were formed by southeastward movement of stratigraphy, and structural traps in the Magdalena
a single regional thrust sheet with minor imbrications. Valley and Llanos foothills.
Initially this sheet moved upward above a ramp
extending from a mid-crustal level to the widespread
Upper Cretaceous shale unit, which it followed south-
eastward as a blind thrust sheet, causing delamination
between the overlying and underlying sections at its Acknowledgments The author gratefully acknowledges
leading edge. The present offset between the two fruitful discussions with A. W. Bally, G. Cathyl-Bickford, J. W.
Cordilleras arose, as did the offset between the Kerr, and J. K. Lentin and constructive criticism by AAPG
Magdalena and Cesar valleys, from the presence of a critical readers. Support for the study came from International
lateral ramp within this thrust sheet. Like the frontal Tectonic Consultants Ltd.
ramps that it connects, the lateral ramp extended
upward from a glide horizon within the basement as far
as the Upper Cretaceous shales of the Umir Formation,
an interval of over 10 km. Movement of the Bucara- REFERENCES CITED
manga fault, mainly vertical, occurred within the
hanging wall of the overthrust sheet in response to the Audemard, F., 1993, Triangle zones of the Maracaibo basin
underlying footwall lateral ramp. Subsequent southeast- (abs.): AAPG Bulletin, v. 77, p. 303.
Banks, C. J., and J. Warburton, 1986, “Passive-roof” duplex
ward movement of the southward-propagating Merida geometry in the frontal structures of the Kirthar and
Andes thrust sheet carried the Perija–Eastern Cordillera Sulaiman mountain belts, Pakistan: Journal of Structural
thrust sheet farther southeastward, producing the Geology, v. 8, p. 229–237.
present configuration of Perija, Merida, and Eastern Bell, J. S., and P. B. Jones, 1993, Triangle zones and under-
Cordillera (Figure 12). thrusting along the southern margin of the Cordillera de la
Westward gravity sliding in the Middle Magdalena Costa, north-central Venezuela (abs.): AAPG Bulletin, v.
Valley began close to or at the end of the deformation 77, p. 306–307.
process, at the junction of the Magdalena Valley and Butler, K., and S. Schamel, 1988, Structure along the eastern
Bucaramanga fault. The gravity slides may be associated margin of the Central Cordillera, Upper Magdalena Valley,
with the latest movement of the Bucaramanga fault. Colombia : Journal of South American Earth Sciences, v.1,
p. 109–120.
Subsequent westward-directed slip sheets occurred Colletta, B., F. Hebrard, J. Letouzey, P. Werner, J.-L. Rudkei-
progressively farther south along the eastern edge of the wicz, 1990, Tectonic style and crustal structure of the
valley, perhaps triggered by southward propagation of Eastern Cordillera (Colombia) from a balanced cross-
the southeast-vergent thrust system underlying the section, in J. Letouzey, ed., Petroleum and tectonics in
Merida Andes. mobile belts: Paris, Editions Technip, p. 81–100.
658 Jones

Dengo, C. A., and M. C. Covey, 1993, Structure of the Eastern Laubscher, H. P., 1987, The kinematic puzzle of the Neogene
Cordillera of Colombia: implications for trap style and Northern Andes, in J.-P. Schaer and J. Rodgers, eds., The
regional tectonics: AAPG Bulletin, v. 77, p. 1315–1337. anatomy of mountain ranges: New Jersey, Princeton
Gordy, P. L., F. R. Frey, and D. K. Norris, 1977, Geological University Press, p. 211–227.
guide for the CSPG 1977 Waterton–Glacier Park field McClay, K. R.,1992, Glossary of thrust tectonic terms, in K. R.
conference: Canadian Society of Petroleum Geologists, McClay, ed., Thrust tectonics: London, Chapman and Hall,
Calgary, 93 p. p. 419–433.
Graham, R. 1981, Gravity sliding in the Maritime Alps, in K. Morley, C. K., 1986, A classification of thrust fronts: AAPG
R. McClay and N. J. Price, eds., Thrust and nappe tectonics: Bulletin, v.70, p. 12–15.
Geological Society of London Special Publication 9, p. Schamel, S., 1991, Middle and Upper Magdalena basins,
335–352. Colombia, in K. T. Biddle, ed., Active margin basins:
Gramond, J.-F., 1993, Normal faulting and tectonic inversion AAPG Memoir 82, p. 283–301.
driven by gravity in a thrusting regime: Journal of Struc- Suppe, J., 1985, Principles of structural geology: New Jersey,
tural Geology, v. 16, p. 1–10. Prentice-Hall, 537 p.
Harrison, J. V., and N. L. Falcon, 1936, Gravity collapse struc- Taborda, B., 1965, The geology of the de Mares concession,
tures and mountain ranges, as exemplified in south- 1965, in Geological field trips, Colombia, 1958–1978:
western Iran: Quarterly Journal of the London Geological Colombian Society of Petroleum Geologists, p. 119–159.
Society, v. 92, p. 91–102. Teal, P. R., 1983, The triangle zone at Cabin Creek, Alberta, in
Ingeominas, 1988, Mapa Geologico de Colombia: Ingeominas, A. W. Bally, ed., Seismic expression of structural styles:
Instituto Nacional de Investigaciones Geologico-Mineras, AAPG Studies in Geology Series 15, v. 3, no. 4.1, p. 48–53.
Bogota, 2 sheets, scale 1:1,500,000. Urbina, C., 1993, Structural style of the southern flank of
Irving, E. M., 1975, Structural evolution of the northernmost Merida Andes, Venezuela (abs.): AAPG Bulletin, v.77,
Andes: USGS Professional Paper 846, 47 p. p. 352.
Jones, P. B., 1982, Oil and gas beneath east-dipping under- Vann, I. R., R. H. Graham, and A. B. Hayward, 1986, The
thrust faults in the Alberta foothills, in R. B. Powers, structure of mountain fronts: Journal of Structural
ed.,Geologic studies of the Cordilleran thrust belt: Rocky Geology, v. 8, p. 215–227.
Mountain Association of Geologists, Denver, p. 61–74. Ward, D. E., R. Goldsmith, A. Jimeno, J. Cruz, H. Restrepo,
Jones, P. B., 1987, Quantitative geometry of thrust and fold and E. Gomez, 1977, Mapa geologico del Cuadrangulo H-
belt structures: AAPG Methods in Exploration Series, v. 6, 12, “Bucaramanga”–Colombia: Instituto Nacional de
26 p. Investigaciones Geologico-Mineras, 1:100,000.
Jones, P. B., 1988, Significance of blind or low-angle normal
faults: AAPG Bulletin, v. 72, p. 203.
Jones, P. B., 1990, Oil and gas beneath east-dipping under-
thrust faults in the Alberta foothills, with author’s post-
script (Chinese translation of Jones, 1982, with additions):
Journal of Xinjiang Petroleum Geology, v. 11, p. 191–198.
Jones, P. B., 1993, Gravity sliding, thrusting, and petroleum Author’s Mailing Address
traps in the Magdalena basins and Cordillera Orientale,
Colombia (abs.): AAPG Bulletin, v. 77, p. 327. Peter B. Jones
Julivert, M., 1970, Cover and basement tectonics in the International Tectonic Consultants Ltd.
Cordillera Orientale of Colombia, South America, and a #700, 665 8th Street S.W.
comparison with some other folded chains: GSA Bulletin, Calgary, Alberta T2P 3K7
v. 32, p. 249–271. Canada
Basin Development and Tectonic History of the
Llanos Basin, Colombia

M. A. Cooper
F. T. Addison
R. Alvarez
A. B. Hayward
S. Howe
A. J. Pulham
A. Taborda
BP Exploration (Colombia) Ltd.
Bogotá, Colombia

Abstract

T he Llanos basin lies east of the Eastern Cordillera in northeastern Colombia. Basin development
commenced with a Triassic–Jurassic synrift megasequence related to the separation of North and South
America in the Caribbean. Basin development continued in the Cretaceous as a back-arc megasequence behind
the Andean subduction zone. Marine deposition was abruptly terminated during the early Maastrichtian due
to final accretion of the Western Cordillera.
The accretion of the Western Cordillera created the pre-Andean foreland basin megasequence
(Paleocene–early Miocene), which covered the Magdalena Valley, Eastern Cordillera, and Llanos basin. This
megasequence is dominated by fluviodeltaic strata. The overlying Andean foreland basin megasequence
commenced with deformation in the Central Cordillera and Magdalena Valley. The Andean foreland basin
megasequence also includes the Guayabo Formation, which is a classic molasse sequence shed from the devel-
oping mountains of the Eastern Cordillera as deformation moved eastward into the Llanos foothills. The defor-
mation in the Llanos foothills is a combination of inversion of preexisting extensional faults and thin-skinned
thrusting.

Resumen

L a Cuenca de la Llanos se encuentra localizada al oriente de la Cordillera Oriental en el nororiente de


Colombia. El desarrollo de la cuenca comenzó con una megasecuencia de “synrift” Triasica-Jurasica
relacionada con la separación de Norte y Suramerica en el Caribe. El desarrollo de la cuenca continuó durante
el Cretácico con una megasecuencia de “back-arc” en frente de la zona de subducción de los Andes. La sedi-
mentación marina terminó abruptamente durante el Maestrichtiano temprano debido a la acreción final de la
Cordillera Occidental.
La acreción de la Cordillera Occidental dió origen a una megasecuencia pre-andina de antepaís (Paleoceno
a Mioceno Inferior) la cual cubrió el Valle del Magdalena, la Cordillera Oriental y la Cuenca de los Llanos. Esta
megasecuencia está dominada por sedimentos fluvio-deltaicos. La megasecuencia andina de antepaís supraya-
cente comenzó con la deformación en la Cordillera Central y el Valle del Magdalena. La megasecuencia andina
de antepaís también incluye la Formación Guayabo la cual es una clásica molasa proveniente de la erosión de
la naciente Cordillera Oriental a medida que la deformación se movió hacia el este en el piedemonte de los
Llanos. La deformación en el piedemonte de los Llanos es una combinación de inversión de las fallas exten-
sionales pre-existentes y cabalgamientos de escamación delgada.

Cooper, M. A., F. T. Addison, R. Alvarez, A. B. Hayward, S. Howe, A. J. Pulham, and 659


A. Taborda, 1995, Basin development and tectonic history of the Llanos basin,
Colombia, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 659–665.
660 Cooper et al.

INTRODUCTION
The physiography of Colombia is dominated by the
Andes mountains in the western half of the country and
by the Amazon-Orinoco basin in the east. The
Colombian Andes are split into three ranges—the
Western, Central, and Eastern Cordilleras—which to the
south merge into a single range in Ecuador. To the east of
the Eastern Cordillera is Los Llanos, an elevated
savannah that is part of the catchment area for the Rio
Orinoco (Figure 1). Major work that has been done on
the stratigraphy, tectonics, and regional tectonic setting
of Colombia include Hettner (1892), Hubach (1957),
Bürgl (1961), Etayo-Serna (1979), Fabre (1983), McCourt
et al. (1984), Pilger (1984), Aspden and McCourt (1986),
Ben Avraham and Nur (1987), Megard (1987), Pardo-
Casas and Molnar (1987), Burke (1988), Butler and
Schamel (1988), and Montgomery (1992).
The major tectonic events that have influenced the
development of the Llanos basin are all closely tied to the
development of the active margin of western South
America. The regional structural evolution is divisible
into eight major events:

1. Triassic–Early Cretaceous—Rift basins developed


as a result of the separation of North and South Figure 1—Map of major tectonic provinces and sutures in
America as the Caribbean opened (~235–130 Ma); Colombia.
this is the synrift megasequence.
2. Barremian–Maastrichtian—A prolonged period of
episodic extension occurred on a series of exten- the frontal fold and thrust belt of the Eastern
sional faults (e.g., the Guaicáramo fault system) Cordillera (10.5 Ma–present day).
along with passive regional subsidence in a back-
arc basin setting (~125–74 Ma); this is the back-arc The aim of this paper is to present a brief review of the
megasequence. basin development, chronostratigraphy, and structural
3. Maastrichtian–early Paleocene—The final event in style of the Llanos basin and Llanos foothills to provide a
the accretion of the Western Cordillera caused regional context for the recently discovered Cusiana
uplift and erosion of the Central Cordillera (~74–65 giant field.
Ma); this is the onset of the pre-Andean foreland
basin megasequence.
4. middle Eocene:—An early compressional deforma-
tion event affected the Magdalena Valley and the REGIONAL STRATIGRAPHIC
western margin of the Eastern Cordillera (~49–42 FRAMEWORK AND BASIN EVOLUTION
Ma) due to an increase in convergence rates of the
Nazca and South American plates (Pardo-Casas The basin stratigraphic model has been developed on
and Molnar, 1987; Daly, 1989). the basis of published data and the log, core, seismic, and
5. late Eocene–late Oligocene—A prolonged period of outcrop data acquired by BP during regional studies of
subsidence and localized normal faulting occurred the Llanos basin and exploration of the Llanos foothills.
in response to flexure of the lithosphere in the The chronostratigraphic summary of the Llanos basin
foreland basin created by the deformation load of (Figure 2) is based on a sequence stratigraphic scheme
the Western and Central Cordilleras (~39–29 Ma). originally developed for the Cusiana field and adjacent
6. late Oligocene–early Miocene—Deformation in the areas of the foreland. This stratigraphic scheme was
Cauca and Magdalena valleys caused continued subsequently extended throughout the Llanos basin by
subsidence in the Llanos basin (~29–16.5 Ma). careful correlation of the well logs and using available
7. middle Miocene—A phase of rapid subsidence biostratigraphic data. Comparison of the sequence
occurred as deformation, uplift, and erosion stratigraphy with the conventional industry lithostrati-
commenced in the Eastern Cordillera and estab- graphic scheme for the Llanos basin is illustrated in
lished the foreland basin depocenter in the Llanos Figure 3.
foothills (~16.5–10.5 Ma); this is the Andean The sedimentary rocks in the Llanos foothills and
foreland basin megasequence. Llanos basin were deposited in a basin that evolved from
8. late Miocene–Recent—The latest phase of compres- a back-arc basin in the Late Cretaceous to a foreland
sion and inversion associated with the formation of basin in the early Tertiary and whose depocenter moved
Basin Development and Tectonic History of the Llanos Basin, Colombia 661

Figure 2—Chronostratigraphic summary diagram for the Llanos basin and foothills based on well and outcrop data. The sea
level curve of Haq et al. (1987) has been adjusted to the BP time scale.

progressively eastward throughout the Tertiary. The resulted in deposition of a series of basal, shallow marine
resulting stratigraphy is a highly punctuated succession and shoreline Cretaceous sandstones, which progres-
of Upper Cretaceous–lower Tertiary strata, representing sively onlapped farther eastward onto the Guyana shield
periodic marginal deposition in major contemporaneous (Figure 2). These sandstones are depositional sequence
depocenters to the west and north. The stratigraphic K50 (Une Formation equivalent) (Hubach, 1931).
record becomes more complete westward in the Eastern In the Turonian–early Coniacian (91–88 Ma), global
Cordillera and the Magdalena Valley, although Tertiary sea level rise (Haq et al., 1987), combined with anoxic
rocks are only sparsely preserved in the Eastern upwelling conditions, resulted in deposition of a succes-
Cordillera because of late Miocene and Pliocene uplift sion of marine mudstones, cherts, and phosphates
and erosion. The following basin evolution model has (Figure 2). These sediments formed an excellent marine
been developed on the basis of available published infor- source rock (sequence K60, Gacheta Formation) (Miller,
mation and by integrating recent BP well, seismic, and 1979) in the Llanos area. This sequence is the equivalent
outcrop data. of prolific source rocks such as the Villeta Shale
Rocks older than the Late Cretaceous include a succes- Formation in the Upper Magdalena Valley (Beltrán and
sion of Paleozoic metamorphic and sedimentary rocks Gallo, 1968) and the La Luna Formation of the Middle
that have only been penetrated in a few wells in the Magdalena Valley and western Venezuela.
Llanos. Triassic–Lower Cretaceous rocks are absent in Sequence K60 deposition was terminated by a fall in
the area except for possible small, localized synrift relative sea level in the Coniacian–early Santonian (88–85
sequences. Upper Cretaceous strata thus generally rest Ma). The Llanos foothills area was on the eastern margin
directly on Paleozoic basement. of the basinal system. Sequences K70 and K80 (which
Upper Cretaceous deposition of the back-arc megase- equate approximately with the Guadalupe; Hettner,
quence was initiated in Cenomanian time (98–91 Ma) 1892) were deposited at this time and represent two
during a regional transgression that drowned the major cycles of eastward shoreline progradation, aggra-
exposed Paleozoic rocks of the Llanos region and dation, and retrogradation. They are dominated by high-
662 Cooper et al.

TYPE LOG SIGNATURE sequence K80 is a shale unit that has been mistakenly
PROPOSED CONVENTIONAL
FORMATION
identified as the Maastrichtian–Paleocene Guaduas
AGE SEQUENCES GAMMA RESISTIVITY
RAY NAMES Formation (Figure 3) (Sarmiento, 1992) in some of the
0m FARALLONES
earlier wells in the Llanos foothills (e.g., Medina-1).
Recently acquired data by BP has conclusively dated
these youngest Cretaceous rocks in the foothills as
Campanian.
PLIOCENE
500 m The final accretion event in the Western Cordillera

UPPER
commenced at the end of the Cretaceous. A relative drop
in sea level, probably linked to the onset of compression
to the west, resulted in a fundamental change in the
T90 nonmarine deposition of the pre-Andean foreland basin

GUAYABO
1000 m

megasequence. Sequence T10 is not present in the Llanos


basin and foothills, being represented by a hiatus of
LATE

about 20 m.y. that spans the Cretaceous–Tertiary

LOWER
boundary.
1500 m
Renewed deposition commenced about 60 Ma in the
late Paleocene in response to a far-reaching transgres-
sion. The Barco Formation (Notestein et al., 1944) forms
the basal transgressive part of sequence T20, which was
2000 m
T80 laid down on a major unconformity surface. It mainly
MIDDLE
MIOCENE

comprises sandstone-rich estuarine deposits. Marine


LEON

influence is strong throughout the Barco Formation in


the Cusiana field area in the Llanos foothills, but at the
2500 m top of the formation there is a relatively abrupt upward
transition into more heterolithic coastal plain and alluvial
plain deposits. T20 sandstone deposition ended as the
T70 C1
late Paleocene transgression weakened and a relative
EARLY

3000 m highstand in sea level was established (~59 Ma). During


C2
CARBONERA

the subsequent regression, the regional shoreline


C3 gradually shifted westward. The sediments laid down
T60 C4
during this sea level highstand were muddy lower
C5 coastal plain deposits (Los Cuervos Formation)
T50 3500 m
(Notestein et al., 1944). In the Llanos, a hiatus of up to
C6 20 m.y. separates sequences T20 and T30 and appears to
OLIGOCENE C7 correlate with a similar unconformity in the Middle
T40 Magdalena Valley.
4000 m
C8 Deposition in the area was renewed in the
LATE
EOCENE T30 MIRADOR middle–late Eocene (~39 Ma). Deposition of sequence
LOS CUERVOS
PALEOCENE
T20 BARCO
T30 (Mirador Formation) (Notestein et al., 1944) began in
GUADUAS response to a far-reaching transgression that came out of
CAMPANIAN K80 4500 m
U. GUADALUPE the foreland basins to the west and north. Initial T30
L. CRETACEOUS

GUADALUPE SHALE
SANTONIAN K70 L. GUADALUPE
deposition included marine-influenced, sand-rich, valley
fill deposits that passed upward into muddier coastal
GACHETA plain sediments. Continued transgression eventually
CONIACIAN-
TURONIAN K60 5000 m SHALES
submerged this middle Mirador alluvial plain and estab-
lished a shallow marine shelf across the Cusiana area.
CENOMANIAN-
ALBIAN
K50 UNE
Offshore muds and sandy bioturbated shoreface progra-
dational cycles punctuated by sand-rich fluvial and
estuarine valley fill deposits comprised latest Eocene
Figure 3—Comparison of stratigraphic schemes for the
Llanos foothills and Llanos basin. The symbols in the deposition, forming the upper part of the Mirador
sequences column are the same as in Figure 5. The Formation. All of the coarser grained sandstones in the
symbols in the stylized resistivity log column indicate the Mirador Formation in the Llanos foothills are extremely
depositional environments and are the same as in Figure 2. mature quartz arenites.
After sequence T30 deposition, four major cycles of
marine-influenced lower coastal plain deposition
energy, quartz-rich shoreface sandstones supplied from occurred in the Llanos basin and foothills (sequences
the Guyana shield to the west and exhibit a widespread T40–T70). These sequences are ~34–16.5 Ma and are
distribution across the Llanos basin and foothills. The traditionally termed the Carbonera Formation (Notestein
Campanian K80 sandstones form the oldest proven et al., 1944). These cycles are bounded by widespread
commercial reservoir unit in the foothills. At the top of maximum flooding surfaces. Each cycle consists of a
Basin Development and Tectonic History of the Llanos Basin, Colombia 663

mud-dominated highstand systems tract followed by a N


ºW 6º
thin, forced retrogradational systems tract, and ending 72
with a sand-prone transgressive systems tract that culmi- Tamara-1
nates with the maximum flooding surface. The
sequences are thus not true sequences in the sense of PAZ DE
Mitchum et al. (1977), but are genetic stratigraphic units ARIPORO

YOPAL FAUL T SYS TEM


in the sense of Galloway (1989). Type 1 sequence bound-
aries (Mitchum et al., 1977) have not yet been recognized.
The units can be correlated throughout the Llanos basin N

and show a gradual increase in the sand percentage as
Pauto-2
the Guyana shield provenance area is approached. RI O P A UTO
In the middle Miocene, the global rise in sea level

M
coincided with the first significant deformation and º

SYSTE
72 'W
uplift in the Eastern Cordillera and hence with a signifi- 30
cant loading event that tectonically enhanced the relative
rise in sea level and the highstand systems tract that

NUNCHI A SYNC LINE


resulted (T80 mudstone, Leon Formation) (Notestein et

F A ULT
RI
O T OCARIA
al., 1944). Evidence for partial emergence of the Eastern
Cordillera is that sequence T80 becomes more sand
prone in the western part of the foothills, suggesting that 72
ºW
there was a supply of coarse clastics derived from the
west.
The final depositional episode in the Llanos was the
El Morro-1

YOPAL
N

GUAICARAMO
deposition of about 3000 m of coarse continental clastics
in sequence T90 (Guayabo Formation) (Hubach, 1957)
from ~10 to 2 Ma ago. This last phase of deposition

YOPAL FAULT
N

marks uplift of the Eastern Cordillera immediately west
of the foothills and migration of the foreland basin axis to
the current location of the Eastern Cordillera foothills
(Figure 2). Deposition of this molasse unit caused rapid Cupiagua-1

late stage burial of the Upper Cretaceous–lower Tertiary


Cusiana-2A
stratigraphic section in the foothills and Llanos basin.
LINE OF SECTION

ºW Cusiana-4
73 RIO CUSIANA
Leticia-1
STRUCTURAL EVOLUTION AND STYLE CUSIA NA AN TICLINE

The boundary between the Eastern Cordillera and CHITAM CUSIANA F A LT


RIO E NA
U
the Llanos foothills is the Guaicaramo fault system
(Figure 4). The foothills are about 15–20 km wide and are
separated from the foreland to the north and south of the 72
º
'W
area by the inverted Cusiana-Yopal fault system. In the 30
N
central part, the inversion faults lie beneath the thin- 5º

skinned Yopal fault system (Figure 4), which detaches in


the lower part of sequence T40. The earliest documented
0 5 10 15 20 km
extension on the Cusiana-Yopal fault system occurred in
the Late Cretaceous during deposition of sequence K80
LIN E

(upper Guadalupe Formation). This is based on the Medina-1


IO ANTIC

differences in thickness of K80 measured in the Cusiana GUAYABO


field wells and wells in the immediate foreland (e.g., LEON
GU A V

Leticia-1). It is likely, however, that the fault trend had an


earlier extensional history during Early Cretaceous CARBONERA
rifting and back-arc subsidence, given the dramatic thick- MIRADOR +
ening of the Lower Cretaceous strata from the foreland CRETACEOUS
toward the Eastern Cordillera (Hebrard, 1985; Ulloa and ANTICLINE
Rodriguez, 1981). Extension continued episodically from SYNCLINE
Late Cretaceous to middle Miocene time (sequence T80, THRUST FAULT
Leon Formation) and was punctuated by hiatuses and
tectonic quiescence. The extensional movements on the
Cusiana fault system can be seen on seismic data and are Figure 4—Surface geologic map of the Llanos foothills
also demonstrable from the thickness of the sequences in showing major structures and the location of the cross
the foreland wells as compared to wells within the section and wells referred to in the text.
664 Cooper et al.

NW SE
Guaicaramo Cusiana-4 Cusiana-2A Leticia-1
Fault System Cusiana
Yopal Fault Fault

0 5 10 Km

Figure 5—Cross section through the Cusiana field in the Llanos foothills. See the “proposed sequences” column in Figure 3
for the key to the stratigraphic units.

hanging wall of the fault. This phase of extension on the CONCLUSIONS


Cusiana fault system accommodates the flexure of the
lithosphere in response to the load imposed by the The tectonic history of the region records an initial
regional compressional deformation. phase of localized Triassic–Jurassic rifting followed by an
As deformation in the Eastern Cordillera migrated Early Cretaceous back-arc basin that became less active
eastward, the foothills became involved in the frontal by the middle Cretaceous. In the Late Cretaceous, the
fold and thrust belt (R. Herrera, 1971, personal commu- collision of the Western Cordillera initiated the foreland
nication; Colletta et al., 1990). Relatively simple basin megasequence that has dominated to the present
compression along a WNW-ESE trending azimuth day.
caused inversion along the Cusiana-Yopal fault system The tectonic evolution strongly influenced strati-
(Figure 5). graphic development of the basin. The back-arc mega-
The thin-skinned Yopal fault, which detaches within sequence is characterized by marine strata that are more
sequence T40, overrides the Cusiana fault to the north basinal to the west. The collision event in the Late Creta-
and buries the branch line with the latter fault. To the ceous reestablished continental and coastal plain envi-
west of these frontal inversion structures is a system of ronments throughout the basin. The pre-Andean
major regional synclines. The western limbs of the foreland basin megasequence shows strongly cyclical
synclines are elevated by a series of structures that sediment patterns with the alternation of continental,
involve the Upper Cretaceous and lower Tertiary sedi- coastal plain, and marginal marine environments. This is
mentary sequences. These structures can be modeled as a interpreted to be the result of the highly peneplained
series of basement involved or thin-skinned (Gacheta Guyana shield hinterland, which had a low paleoslope
Shale, K60 detachment) duplex horses (Figures 4, 5), gradient and was thus susceptible to rapid flooding and
largely based on the evidence of fault repetition of the regression. This in turn allowed the rapid migration of
Mirador in the El Morro-1 well. facies belts within the systems tracts. The main Andean
The recent paper by Dengo and Covey (1993) contains deformation phase caused inversion of preexisting exten-
a cross section through the Unete-1 well that implies a sional faults and thin-skinned thrust structures.
deeper detachment for the Yopal fault than is evident
from our studies of the foothills (Figure 5), which are
based on an extensive seismic and well database. The Acknowledgments We wish to thank BP Exploration
other difference in the work by Dengo and Covey (1993) (Colombia) Ltd. for permission to publish this work, Michel
is the absence of inversion on the basement-involved Coudèyre of Total for the type log signatures used, and Andres
Cusiana-Yopal fault system. Tovar for drafting the figures.
Basin Development and Tectonic History of the Llanos Basin, Colombia 665

REFERENCES CITED McCourt, W. J., J. A. Aspden, and M. Brook, 1984, New


geological and geochronological data from the Colombian
Aspden, J. A., and W. J. McCourt, 1986, Mesozoic oceanic Andes: continental growth by multiple accretion: Journal
terrane in the central Andes of Colombia: Geology, v. 14, of the Geological Society of London, v. 141, p. 831–845.
p. 415–418. Megard, F., 1987, Cordillera Andes and Marginal Andes: a
Beltrán, N., and J. Gallo, 1968, The geology of the Neiva sub- review of Andean geology north of the Arica Elbow (18
basin, Upper Magdalena basin (southern portion): Ninth deg. S), in J. W. H. Monger and J. Francheteau, ed.,
Annual Field Conference Guidebook, Colombian Associa- Circum-Pacific orogenic belts and evolution of the Pacific
tion of Petroleum Geologists and Geophysicists, Bogotá, Ocean basin: American Geophysical Union, Geodynamics
p. 253–274. Series, v. 18, p. 71–95.
Ben-Avraham, Z., and A. Nur, 1987, Effects of collisions at Miller, T., 1979, The geology of the Eastern Cordillera between
trenches on oceanic ridges and passive margins, in J. W. H. Aguazul–Sogamoso–Villa de Leiva, part I, in Geological
Monger and J. Francheteau, Circum-Pacific orogenic belts Field-Trips, Colombia 1959–1978: Colombian Society of
and evolution of the Pacific Ocean basin: American Petroleum Geolologists and Geophysicists, Geotec Ltda,
Geophysical Union, Geodynamics Series, v. 18, p. 9–18. Bogotá, p. 349–396.
Bürgl, H., 1961, Sedimentación cíclica en el geosinclinal Mitchum, R. M., P. R. Vail and S. Thompson III, 1977, Seismic
Cretáceo de la Cordillera Oriental de Colombia: Servicio stratigraphy and global changes of sea level, part 2: the
Geológico Nacional, Informe No. 1347, 60 p. depositional sequence as a basic unit for stratigraphic
Burke, K., 1988, Tectonic evolution of the Caribbean: Annual analysis, in C. E. Payton, ed., Seismic stratigraphy—appli-
Review of Earth and Planetary Sciences, v. 16, p. 201–230. cations to hydrocarbon exploration: AAPG Memoir 26,
Butler, K., and S. Schamel, 1988, Structure along the eastern p. 53–62.
margin of the Central Cordillera, Upper Magdalena Valley, Montgomery,S., 1992, Petroleum potential of Upper and
Colombia: Journal of South American Earth Sciences, v.1, Middle Magdalena basins, Colombia, part 2: plate
p. 109–120. tectonics, reservoirs, source rocks, and field histories:
Colletta, B., F. Hebrard, J. Letouzey, P. Werner, and J.-L. Petroleum Frontiers, v. 9, 67 p.
Rudkiewicz, 1990, Tectonic style and crustal structure of Notestein, F. B., C. W. Hubman, and J. W. Bowler, 1944,
the Eastern Cordillera (Colombia) from a balanced cross- Geology of the Barco concession, Republic of Colombia:
section, in J. Letouzey, ed., Petroleum and tectonics in GSA Bulletin, v.55, p. 1155–1218.
mobile belts: Paris, Editions Technip, p. 81–100. Pardo-Casas, F., and P. Molnar, 1987, Relative motion of the
Daly, M. C., 1989, Correlations between Nazca/Farallon plate Nazca (Farallon) and South American plates since Late
kinematics and forearc basin evolution in Ecuador: Cretaceous time: Tectonics, v. 6, p. 233–248.
Tectonics, v. 8, p. 769–705. Pilger, R. H., Jr., 1984, Cenozoic plate kinematics subduction
Dengo, C. A., and M. C. Covey, 1993, Structure of the Eastern and magmatism: South American Andes: Journal of the
Cordillera of Colombia: implications for trap styles and Geological Society of London, v. 141, p. 793–802.
regional tectonics: AAPG Bulletin, v. 77, p. 1315–1337. Sarmiento, G., 1992, Estratigrafía y medios de depósito de la
Etayo-Serna, F, 1979, Zonation of the Cretaceous of Central Formación Guaduas: Ingeominas, Bogotá, Boletín
Colombia by ammonites: Publicacion Especial Ingeominas, Geológico, no. 32-1, p. 3–44.
no. 2,p. 1–186. Ulloa, C., and E. Rodríguez, 1981, Geología del cuadrángulo
Fabre, A., 1983, La subsidencia de la Cuenca del Cocuy K-13: Tauramena, Ingeominas, Bogotá, Boletín Geológico,
(Cordillera Oriental de Colombia) durante el Cretáceo y el no. 24-2, p. 3–30.
Terciario, Segunda parte: Esquema de Evolución
Tectónica: Geología Norandina, v. 8, p. 49–61.
Galloway, W. E., 1989, Genetic stratigraphic sequences in
basin analysis, I: architecture and genesis of flooding-
surface bounded depositional units: AAPG Bulletin, v. 73,
p. 125–142. Authors’ Mailing Addresses
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of M. A. Cooper
fluctuating sea levels since the Triassic: Science, v.235, PanCanadian Petroleum Ltd.
p. 1156–1166. 150 9th Ave S.W.
Hebrard, F., 1985. Les foot-hills de la Cordillère Orientale de
P.O. Box 2850
Colombie entre les rios Casanare et Cusiana. Evolution
géodynamique depuis l’Eo Crétacé: Thèse doctorale 3ème Calgary, Alberta T2P 2S5
cycle, Université Pierre et Marie Curie, Paris, no. 85-08, Canada
162 p.
Hettner, A., 1892, Die kordillere von Bogota: Ergzh zu Peter- F. T. Addison
manns Mitteilungen Bd. 22, Erganzungsheft 104, p.1–131. R. Alvarez
Hubach, E., 1931, Geología petrolífera del departamento de A. B. Hayward
Norte de Santander: Servicio Geológico Nacional, Bogotá, S. Howe
Informe 176, parte A, p. 1–218, parte B p. 219–416, parte C A. J. Pulham
26 láminas. A. Taborda
Hubach, E., 1957, Contribución a las unidades estratigráficas
BP Exploration (Colombia) Ltd.
de Colombia, (enumeración regional, de mas reciente a
mas antiguas): Servicio Geológico Nacional, Informe no. Carrera 9A, 99-02, Bogotá
1212, 165 p. Colombia
Crustal Architecture and Strain Partitioning in the
Eastern Venezuelan Ranges

H. Passalacqua Y. Gou
F. Fernandez Beicip
Cedex, France
Intevep, S.A.
Caracas, Venezuela F. Roure
IFP
Cedex, France

Abstract

T he eastern Venezuelan Coast Ranges result from oblique convergence along the South
American–Caribbean plate boundary, expressed at the surface by the El Pilar dextral strike-slip fault. A
crustal-scale, balanced NNW-SSE cross section has been constructed across this major transfer zone that links
oceanic subduction of the Lesser Antilles with continental subduction of the Andes. It shows a major discrep-
ancy between the cover and basement lengths, which can be explained by tectonic inheritance from the
Tethyan margin and an initially thinned crust and basement tilted blocks. The section, interpreted down to the
Moho, is constrained by magnetic and gravimetric profiles. A major gravimetric low along the axis of the
Maturín basin shows the progressive northward deepening of the Moho. Positive magnetic anomalies on the
southern flank of this basin probably result from shallow basaltic intrusions along the thinned part of the pale-
omargin or from crustal heterogeneities. A high-density intracrustal wedge is needed to fit the gravimetric
high north of the Serranía; the solution requires a deep crustal root beneath the belt and a northward-dipping
South American Moho. The results of successive gravimetric and magnetic modeling studies are compared
with the present-day seismicity of the South American–Caribbean plate boundary. Shallow seismicity is
restricted to the eastern part of the El Pilar fault, whereas deep focal mechanisms pick out a northwestward-
dipping subduction slab off the northeastern Venezuelan and Trinidad coasts.
A consistent geodynamic model involving northward-dipping subduction of at least 70 km of South
American continental lithosphere is thus proposed for the area. The El Pilar fault is a shallow structure that
branches at depth on an intracrustal backstop and at the surface transfers the lateral motion required to balance
the northward-dipping subduction in an oblique convergence regime. Petroleum generation was initiated in
the early stages of the flexural evolution of the eastern Venezuelan basin, and early generated oils migrated
over long distances from the inner parts of the belt toward the Orinoco to fill the traps of the Faja Petrolífera.
Tectonic overburial also induced a late stage petroleum generation, with shorter migration of oil toward the
frontal structures (the El Furrial and Orocual trends).

Resumen

L a Serranía del Interior resulta de la convergencia oblicua a lo largo del borde de las placas Suramericana
y del Caribe, que se expresa en superficie por la Falla de El Pilar, de rumbo deslizante dextral. Para este
trabajo se construyó una sección balanceada hasta la corteza a partir de información de superficie y subsuelo, a
través de esta zona de transferencia que une la subducción oceánica de las Antillas Menores con la subducción
continental de Los Andes. La sección geológica en cuestión, se extiende desde la Falla de El Pilar hasta el Rio
Orinoco con rumbo NNW-SSE, cruzando el cinturón de cabalgamientos de la Serranía del Interior, volcados
hacia el sur y la Cuenca de antepaís asociada, la cuenca de Maturín. La geometría actual de esta sección
muestra una discrepancia notable entre las longitudes del basamento y de la cobertura sedimentaria, que se
puede explicar por una herencia tectónica del margen del mar de Tethys y por una corteza inicialmente
delgada y bloques de basamento basculados. La sección geológica, interpretada hasta el Moho, fue controlada
por perfiles magnéticos (transformados al polo) y gravimétricos. El mapa de anomalía de Bouguer muestra un
mínimo de grandes proporciones que se extiende a lo largo del eje de la Cuenca de Maturín, delineando la
profundización progresiva del Moho por debajo de la zona de estudio. El flanco sur de esta Cuenca está carac-
terizado por anomalías magnéticas producidas probablemente por intrusiones basálticas muy someras en la

Passalacqua, H., F. Fernandez, Y. Gou, and F. Roure, 1995, Crustal architecture and strain 667
partitioning in the eastern Venezuelan Ranges, in A. J. Tankard, R. Suárez S., and H. J.
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 667–679.
668 Passalacqua et al.

porción más delgada del paleo-margen o por heterogeneidades en la corteza. Para modelar el alto gravimétrico
al norte de la Serranía fue necesario introducir una cuña de alta densidad; esta solución requiere una raíz
profunda en la corteza por debajo del cinturón de cabalgamientos y el Moho de la Placa Suramericana profun-
dizando hacia el Norte. Los resultados de la sucesivas aproximaciones del modelaje gravimétrico y magnético
se comparan con la sismicidad actual del borde de placas Suramerica-Caribe. En el sector Este de la Falla de El
Pilar la sismicidad es superficial, mientras que mecanismos focales profundos sugieren la presencia de una
placa que subduce al NE de las costas de Venezuela y Trinidad.
Se propone entonces un modelo geodinámico consistente cuya característica principal es una subducción
hacia el Norte de al menos 70 km de litósfera continental de la placa Suramericana (principalmente manto intra
continental y partes de la corteza inferior). La Falla de El Pilar es una estructura superficial que se bifurca a
profundidad en una cuña intracortical y en la superficie transfiere el movimiento lateral necesario para
balancear la subducción hacia el Norte en un régimen de convergencia oblicua. La generación de hidrocarburos
comenzó en las etapas tempranas de la evolución flexural de la Cuenca Oriental de Venezuela, y esos hidrocar-
buros migraron largas distancias hasta la Faja Petrolífera del Orinoco. La sobrecarga tectónica indujo también
una segunda generación, con migraciones mas cortas hacia las estructuras del frente de deformación (alinea-
ciones de El Furrial y Orocual).

INTRODUCTION GEOLOGIC AND GEOPHYSICAL


Oblique convergence is a common phenomenon BACKGROUND
along plate boundaries and usually results at the surface
level in the juxtaposition of a major strike-slip Dynamics of Eastern Venezuela
(transform) fault and a doubly warped (V-shaped)
orogenic belt. This is the case in California where the The Venezuelan Coast Ranges outline a major transfer
active San Andreas dextral fault is closely associated zone between the westward-dipping B-subduction of the
with the Transverse Ranges (Sylvester and Smith, 1976; Atlantic oceanic lithosphere beneath the Lesser Antilles
Wentworth and Zoback, 1989), in the Pyrenees with the volcanic arc and the westward-dipping A-subduction of
North Pyrenean sinistral fault (Choukroune et al., 1989; the South American continental lithosphere beneath the
Roure et al., 1989), and in the Alps with the Insubric line Andes (Figure 1) (Weeks et al., 1971; Pérez and
(Laubscher, 1988; Roure et al., 1990). Deep seismic Aggarwal, 1981; Kellog and Bonini, 1982; Stein et al.,
profiles have recently been recorded in such terranes 1982; Wadge and Burke, 1983; Schubert, 1984; Burke et
(Lemiski and Brown, 1988; Choukroune et al., 1989; al., 1984; Beck, 1986; Westbrook and McCann, 1986). At
Roure et al., 1990) and show the deformation of the conti- the surface, its eastern part is characterized by the occur-
nental lithosphere at depth along the plate boundary. rence of a major transform fault, the dextral El Pilar fault
Similarly, analogous experiments have been attempted (Schubert, 1979; Vierbuchen, 1984) (Figures 1, 2).
to simulate the effects of oblique convergence on the A south- to southeast-verging fold and thrust belt is
architecture of fold and thrust belts and to help under- developed south of the El Pilar (Serranía del Interior),
stand how a constant balance is maintained between and a related Neogene flexural basin (Maturín basin)
frontal or deep accretion and superficial strike-slip separates the Mesozoic surface outcrops of the Vene-
motion (Soula, 1984; Malavielle and Cobb, 1986). These zuelan Coast Ranges from the basement of the Guyana
concepts can be applied to other orogenic belts where shield, which crops out south of the Orinoco River
deep seismic profiles have yet to be recorded. For (Figure 1) (Creole Petroleum Corporation, 1965; Potié,
example, the eastern Venezuelan Coast Ranges result 1989). Northwest-southeast oriented strike-slip faults
from oblique convergence at the South American– (from west to east, the Urica, San Francisco, and Los
Caribbean plate boundary. This in turn resulted in the Bajos faults) fragment the Serranía into various blocks
development of the El Pilar dextral strike-slip fault in the (Wilson, 1968; Rosales, 1972; Munro and Smith, 1984).
north and the Maturín basin in the south (Maresch, 1974; Southwest-northeast (Pirital) or WSW-ENE (El Furrial
Ladd, 1976; Schubert, 1981; Beck, 1984; Speed, 1985; and Orocual) oriented thrust fronts limit the extent of the
Stephan et al., 1986; Pindell and Barrett, 1988). allochthon to the south (Figures 1, 2, 3). A detailed
The object of this paper is to propose a consistent description of this thrust system is given elsewhere
geodynamic model for the area. To address this (Roure et al., 1994).
objective, a crustal-scale balanced cross section was North of the El Pilar fault, outcrops are restricted
compiled from available surface and subsurface data in onshore to the Paria and Araya Península and offshore to
the eastern Venezuelan Coast Ranges and was Margarita island (Figure 2). They show either highly
constrained by magnetic and gravimetric profiles. The metamorphosed rocks of sedimentary origin or ophiolitic
results of these successive models were compared with rocks derived from the Late Cretaceous–Paleogene
the present-day seismicity of the South American– partial closure of the Tethyan ocean. They were affected
Caribbean plate boundary. by the early deformation of the South American passive
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 669

Figure 1—Location map showing tectonic setting of northern Venezuela. Box shows study area.

Figure 2—Tectonic setting of eastern Venezuela and present seismicity. SFF, San Francisco fault. (Modified after Fernández
and Pérez, 1974; Burke et al., 1984; Speed, 1985; Bouysse and Westercamp, 1990.)
670 Passalacqua et al.

Figure 3—Geologic map of the Serranía (simplified after Creole Petroleum Corporation, 1965). A–A' locates the modeled
section, and B–B' indicates the part imaged by seismic.

margin (Vignali, 1979; Wadge and MacDonald, 1985; Calc-alkalic volcanic activity related to oceanic
Chevalier et al., 1985, 1988) (Figure 3). Recent offshore subduction is well expressed in the Lesser Antilles, but
extensional structures are superposed on older compres- progressively dies out southwestward, the last morpho-
sive structures between the mainland and Margarita logic expression of the arc being found on Margarita
Island (Case, 1982; Schubert, 1982; Biju-Duval et al., 1983; Island. North of Margarita, an extending back-arc basin
Robertson and Burke, 1989). is developed (the Grenada basin) which separates the
The El Pilar fault and its western extension, the San still active Antilles arc from a fossil arc (the Aves Ridge)
Sebastián fault (Figure 1), connect with the Lesser (Figure 1) (Biju-Duval et al., 1978; Bouysse, 1988).
Antilles frontal thrust to the east and with the Boconó
fault in the Venezuelan Andes to the west. Well Major Tectonostratigraphic Units and
expressed at the surface by the morphologic depression Terranes of the Eastern Venezuelan Ranges
of the Cariaco and Paria gulfs, the present activity of the
El Pilar fault is also reflected in hydrothermal activity South of the El Pilar fault, the recognized Mesozoic
and in seismicity with focal depths shallower than 12 km allochthonous units of the Serranía del Interior, as well as
(Fernández and Pérez, 1974; Kafka and Weidner, 1981; the substratum of the Maturín foreland basin, are
Wadge and Hudson, 1986; Giraldo, 1990). composed of Cretaceous–Paleogene sequences of the old
Elsewhere, recent shallow seismicity is restricted to Tethyan passive margin (Burke, 1988; Eva et al., 1989;
the Los Bajos fault which separates the Serranía del Fourcade et al., 1991). The Mesozoic sequence exposed in
Interior from Trinidad in the east. South of the El Pilar the Serranía and extensively drilled in the foreland rests
fault and west of Trinidad (in the study area), only directly on top of Precambrian basement to the south
diffuse seismicity is presently observed. Deep focal near the Orinoco River, but is detached from its pre-
mechanisms related to the Lesser Antilles arc extend Cretaceous substratum in the north. It becomes progres-
northwestward off the Trinidad and eastern Venezuelan sively thicker northward, reaching a thickness of 3 km in
coasts (Figure 2) (Wadge and Shepherd, 1984; Speed, the allochthon (Potié, 1989).
1985; Westbrook and McCann, 1986; Bouysse and Although they never crop out and have not yet been
Westercamp, 1990). drilled in the area, sedimentary rocks of Paleozoic–
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 671

Figure 4—Seismic profile of


the southern part of the
transect showing the Pirital
high and Morichito basin,
which separate the Maturín
basin from the outcrops of
the Serranía. See Figures 3
and 5 for location.

Jurassic age are inferred on the basis of seismic data to Late Cretaceous or Paleogene time and have probably
extend at depth beneath the Maturín basin, northward traveled far along the transform Caribbean–South
to the Serranía del Interior, and beneath either evapor- American plate boundary (Sykes et al., 1982; Dewey and
ites or coal measures along which the remaining sedi- Pindell, 1985; Erlich and Barrett, 1990). In fact, the first
mentary cover has been detached (Roure et al., 1994) foredeep deposits of both the Maturín basin and the
(Figures 4, 5). Serranía are no older than late Oligocene or early
The Maturín foreland basin is filled with synorogenic Miocene, attesting to the Neogene age of the tectonic
terrigenous deposits, mainly deep water turbidites of the inversion of this part of the margin. This suggests an
lower Carapita Formation (lower–middle Miocene) and allochthonous origin for the older metamorphic rocks
nearshore to continental facies of the upper Carapita, La north of the El Pilar. Since deformation of the margin and
Pica, Morichito, Las Piedras, and Mesa formations closure of the Tethyan ocean is supposed to have
(upper Miocene–Plio-Quaternary), which reach a occurred earlier in the west, all these suspect terranes
thickness of 5 km in the axis of the basin (Figure 5). propably originated from a single ancient orogen that
North of the El Pilar fault, various tectonostratigraphic was first emplaced in the west before being dissected
units (terranes) have been recognized, some of them and remobilized during the oblique convergence regime
deriving from the South American paleomargin and of Neogene age.
others directly from the ancient Tethyan ocean (Upper Only relatively young deposits (late Pleistocene Mesa
Jurassic ophiolite from Margarita Island). Unlike the Formation) escaped the compressive deformation, which
autochthon of the Maturín basin or the allochthonous is now active only in eastern Trinidad (Kugler, 1953;
units of the Serranía, these rocks were metamorphosed in Salvador and Stainforth, 1968; Persool, 1990).
672 Passalacqua et al.

Figure 5—Geologic cross


section interpreted on the
basis of seismic and well
data from Maturín basin,
Pirital high, and Morichito
basin and on field surveys
from Serranía.

Geologic and Geophysical Data Used to north of the Serranía trend. Southward, a more diffuse
Constrain the Models positive anomaly is found along the northern border of
the Guyana shield and may reflect a flexural bulge
In addition to published geologic maps of the Serranía (Figure 7). The magnetic profile is also characterized by
and numerous exploration wells in the Maturín basin, a anomalies on the southern flank of the Maturín basin.
large number of conventional seismic profiles have been
recorded between the outcrops of the Serranía del
Interior and the Orinoco River, allowing good calibration
of the geologic section across the area. Using this CRUSTAL-SCALE, BALANCED CROSS
excellent set of data, we were able to construct a regional SECTION FROM EL PILAR FAULT TO
depth section between the Caribbean coast and the ORINOCO RIVER
Guyana shield, across the El Pilar fault and related struc-
tures. The line discussed here is located in the Bergantín Transport Direction and Orientation of the
block, between the Urica and San Francisco strike-slip Section
faults, and runs across Mount Turimiquire, the highest
peak of the Serranía (Figures 3, 4, 5). Its southern part, Far-traveled “suspect” terranes are found north of the
from the Orinoco to the Morichito basin, is well El Pilar fault, but they cannot be restored to their initial
constrained by seismic lines (Figure 4), whereas its positions that were out of the section. Nonetheless, even
northern part across the Serranía del Interior has been if the area has been affected by oblique convergence,
carefully investigated by field surveys (Figure 5). Using most of the deformation observed south of the El Pilar
Locace software (Moretti and Larrére, 1989), we have fault attests to a NNW-SSE direction of transport for the
balanced and restored this cross section to one possible emplacement of the allochthonous units. In the Maturín
preorogenic configuration (Figure 6). basin, the El Furrial and Orocual frontal thrusts effec-
Since deep reflection or refraction seismic profiles are tively trend WSW-ENE. In the Serranía, the Bergantín
not yet available for the area, we decided to better and Caripe blocks have probably been just slightly
constrain the deep interpretations by using gravimetric rotated counterclockwise during thrust emplacement, as
and magnetic modeling, especially for those parts of the indicated by the dextral motion recorded by the Urica
sections that are not imaged by conventional seismic and San Francisco faults and by the general southwest-
profiles. This includes the regional attitude of the Moho northeast trend of the Pirital out-of-sequence thrust
and the first-order structure of the Serranía del Interior. (Roure et al., 1994) and related folds. Unfortunately, no
Gravimetric and magnetic surveys have been recorded paleomagnetic data are yet available to quantify this
in the past, resulting in good quality regional maps, lateral displacement.
which are available in numerical data banks (Figures 7, The modeled section is northwest-southeast trending
8). These files were used here to compile gravimetric and and parallel to the main transport direction, except north
magnetic curves along the western profile (Figure 9), of the El Pilar fault where cross section balancing was not
which were then used to constrain the modeling and fit attempted. Due to the slight rotation observed for the
the calculated curves with the measured ones. Bergantín and Caripe blocks, a certain amount of lateral
The most obvious feature on the gravimetric map is motion may have occurred along the Pirital thrust or the
the occurrence of a regional anomaly doublet: an San Francisco fault, but it is probably of small amplitude
important low on the axis of the Maturín basin is and has not been taken into account in the present
observed, directly adjacent to a gravity high imaged restorations.
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 673

Figure 6—(A) Crustal-scale balanced cross section from the El Pilar fault to the Guyana shield (present geometry). See
Figure 3 for location. A seismic profile of the southern part of the cross section is shown in Figure 4, and a more detailed
geologic interpretation is given in Figure 5. (B) Crustal-scale balanced cross section of initial geometry.

Figure 7—Gravimetric map showing the most relevant tectonic elements. A–A' is the modeled section.
674 Passalacqua et al.

Figure 8—Magnetic map of


the northern part of the study
area (total intensity, reduced
to the pole).

Shortening of the Sedimentary Cover and ducing early (pre-Cretaceous) normal faults that are
Basement Involvement sealed by the Cretaceous. In these new inital geometries,
the thickness of Upper Jurassic synrift sedimentary rocks
The change observed in the seismic image below 2–3 changes laterally between two successive faults.
sec (two-way time) in the Pirital allochthon has been Therefore, most compressive structures involving the
interpreted as indicating pre-Cretaceous basement basement probably reactivate Late Jurassic half-grabens,
involvement along this out-of-sequence thrust, which is and as a result, the actual amount of basement short-
bounded on the west side by a transfer zone, the Urica ening is given by adding the amount of the Late Jurassic
fault (Roure et al., 1994). Similarly, we have assumed that extension to the 45 km first measured.
the San Francisco fault, with a surface geometry similar
to that of the Urica fault, also involved pre-Cretaceous Additional Causes of the Missing
strata at depth and can be divided into two segments: in Basement
the west, a northwest-southeast trending segment acting
as a tear fault with a dextral-normal throw and, in the We increased the amount of basement shortening by
center, an ENE-WSW trending segment accounting for a taking into account the probable occurrence of Late
southward-directed thrust front with only minimum Jurassic normal faults that involve basement. Instead, we
transverse displacement. could have reduced the amount of cover shortening by
By analogy with the younger Pirital thrust, we considering the effects of possible early gliding and
consider that other out-of-sequence thrusts in the Serranía widening of the Cretaceous–Paleogene sedimentary
could also be basement-involved structures. In our initial cover in a precompressive stage, using the seismically
sections, we assumed that basement and Paleozoic– interpreted salt or evaporitic sole as an extensional (or
Jurassic sedimentary sequences were essentially gravitational) detachment level. Such a salt-related
involved in the shortening along the Pirital thrust with a geometry is common along passive margins where evap-
north-south offset greater than 20 km, but also with a orites occur, such as in West Africa (Gabón) and the Gulf
smaller offset along the San Francisco fault and other less of Mexico (Wu et al., 1990).
well expressed structures. The amount of basement and cover shortening was
In both profiles, initial sections and restorations similar along the section, considering that the South
accounted for about 45 km of shortening in the basement American continental crust was thinned northward
and 90 km in the sedimentary cover. We later tried to during the Late Jurassic rifting event. Two contrasting
reduce the discrepancy observed beween basement and solutions could explain the amount of shortening: 45 km
cover shortening by modifying the Cretaceous cut-offs when neglecting precompressive basement extension
beneath the thrusts and by thickening the sedimentary and 90 km when neglecting any precompression gravity
sequence in the initial sections. Nonetheless, we still had gliding of the Cretaceous–Paleogene cover. A median
a big gap on the restored sections between the length of value of 70 km is more realistic. In all solutions, however,
the Cretaceous cover and the length of the Paleozoic or the present length of the section (and of the South
basement. This anomaly can, in fact, be related to the American Moho) is quite different from the initial one
precompressive history of the margin and can be (about 70 km for the median solution). This gap can be
corrected by taking into account the probable occurrence explained by an equal amount of northward-dipping
of extensional structures linked to a Late Jurassic rifting subduction of parts of the South American lithosphere,
event. including the infracontinental upper mantle and possibly
We finally reduced the discrepancy between the even parts of the lower crust (the remaining parts being
amounts of basement and cover shortening by intro- accreted or stacked in a deep crustal orogenic wedge).
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 675

Table 1—Densities and Susceptibilities Used for Modeling

Density Susceptibility
Layer (g/cm3) (cgs 106)
Foredeep 2.25 0
Passive margin 2.50–2.55 0
Rift-prerift deposition 2.65 100
Ophiolite 2.75 300
Metamorphics 2.75 300
Upper crust 2.70–2.80 0–3500
Lower crust 2.95 2000
High-density wedge 3.00 1500

margin evolved in an extensional regime. Therefore, in


addition to the deepening Moho, we also introduced in
our model a high-density wedge to simulate the
backstop that is known to occur in every accretionary
wedge, which should help to compensate for the
Figure 9—(A) Geologic model along profile A–A'. Note the deepening of the mantle.
intracrustal indenter (backstop) and the northward-dipping For gravimetric modeling, specific densities were
Moho. (B) Magnetic model. (C) Gravity model. For B and C, attributed to each significant lithology: the mantle, the
the measured curve is in black and the calculated curve is lower crust, the middle and upper crust (crystalline
dotted. basement), the Paleozoic and Jurassic sedimentary
substratum, the Cretaceous–Paleogene sedimentary
cover, and the Neogene terrigenous sequences (Table 1).
GRAVIMETRIC AND MAGNETIC However, we changed the initial section substantially to
MODELING fit reasonably with the calculated and measured gravi-
metric curves (Figure 9)—especially underneath the
Numerous gravimetric and magnetotelluric surveys Serranía—by modifying the shape and size of the high-
have been conducted locally (Wong et al., 1982; density body, as well as the thickness of the pre-Creta-
Fernández and Passalacqua, 1990; Martín and Espinoza, ceous sedimentary rocks.
1990; Orihuela and Franklin, 1990; Passalacqua et al., The introduction of magnetic modeling (curves
1990). However, only a few modeling studies have been rotated to the pole) provided additional constraints to the
attempted to constrain the deep geometry of the belt. model. High susceptibilities (Table 1) have been postu-
These previous models (Bonini, 1978; Potié, 1989) tried to lated in the upper crust on the southern side of the
explain the gravimetric low observed on Bouguer Maturín basin, where the crust was already significantly
anomaly maps by a crustal root located beneath the axis thinned during the rifting episodes. Basaltic intrusions or
of the Maturín basin and the gravimetric high north of sills have been encountered in some exploration wells in
the Serranía by a relatively shallow Moho in this area. the nearby Guárico basin. Additional heterogeneities in
Nevertheless, these previous models are inconsistent the crust could also be the cause of local anomalies, since
with the geodynamic evolution of the flexural basins greenstones or amphibolites crop out locally in the
recorded by deep seismic profiles around the world. In Guyana shield. By means of these various hypotheses,
most cases, the crustal root is located beneath the uplifted we obtained a good fit between the observed and calcu-
areas (which here is the Serranía), and the foreland basin lated curves for both gravity and magnetism.
related to the bending of the lithosphere (due to the
loading effects of the allochthon) actually corresponds to
a progressively deepening Moho (Watts and Talwani,
1974; Rey et al., 1990; Lillie, 1991). Therefore, in the initial GEODYNAMIC MODEL FOR THE
geologic model used here, we have inferred both a EASTERN VENEZUELAN RANGES
flexure of the Moho and a crustal root beneath the
Serranía. Continental Subduction and Oblique
Potié (1989) has also proposed a model with a Convergence
northward-dipping Moho in which a shallow ophiolitic
body was assumed in the section beneath the Cretaceous The present seismic activity of the area provides
sedimentary cover south of the El Pilar fault in order to evidence for the occurrence of a northwestward-dipping
balance the high positive gravimetric anomaly in the subducted slab off Trinidad and the eastern Venezuelan
Serranía. However, this hypothesis is unlikely. There is coast. Laterally connected to the oceanic subduction of
no way to obduct any ophiolite here before deposition of the Lesser Antilles, this slab may be partially composed
the Cretaceous strata, at a time when the South American of continental lithosphere since part of the allochthonous
676 Passalacqua et al.

units found in the fold and thrust belt on Trinidad are


fragments of the ancient passive margin of the South
American craton and are therefore of continental origin.
The present-day precise geometry of the continent–ocean
boundary between the Atlantic ocean and the South
American craton is still under discussion due to the thick
overlying strata of the Orinoco submarine fan (Speed,
1985; Burke, 1988). This important limit definitely lies
east of the Venezuelan border (Figures 1, 2). According to
our restorations and modeling studies, the same kind of
continental subduction must have been active to the west
during the Neogene. The last compressive event
recorded in the Serranía-Maturín transects was sealed by
the upper Pleistocene Mesa Formation.
In view of our balanced sections, the minimum length
of the South American subducted continental slab must
have reached 70 km. The slab itself was probably more Figure 10—Strain partitioning and geodynamic model of
extensive since this continental subduction followed an the eastern Venezuelan Coast Ranges. Strike-slip motion is
earlier oceanic one that was necessary to consume most restricted to surface areas located north of the El Pilar fault
of the ancient Tethyan ocean which bordered the South and to the intracrustal indenter. In contrast, only north-
American continental margin to the north (with only south shortening occurs south of the El Pilar in the
allochthon, whereas northward-dipping subduction affects
some small obducted fragments now preserved along the distal part of the South American lithosphere.
the El Pilar fault) (Fourcade et al., 1991)

Relationships Between the El Pilar of the fault. At depth, strike slip motion is restricted to
Strike-Slip Fault and Deep Structure the backstop (intracrustal indenter), whereas northward-
dipping subduction affects the underlying remnants of
Present seismicity indicates only shallow focal mecha- South American lithosphere (Figure 10).
nisms along the El Pilar fault and near northwest- However, this Venezuelan example of strain parti-
southeast trending strike-slip faults such as the Los Bajos tioning is not unique. Analog examples of oblique
fault (Figure 2). Similarly, both paleomagnetism and convergence (Malavielle and Cobb, 1986) attest to the
transport direction markers show a western origin for common connection between shallow normal or strike-
the tectonic units found north of the fault, whereas all the slip faults and deeper thrust planes whose merging front
allochthons of the Serranía farther south result from a remains almost parallel to the shallow structures. In
slight southward displacement of initial fragments of this southeastern Asia (Fitch, 1970) and in the Western
part of the South American margin. Canadian Cordillera (Mattauer and Collot, 1986; Van den
In the Pyrenees and the Alps, the North Pyrenean Driessche, 1986), oblique motion along the plate boun-
fault and the Insubric line, respectively, are strike-slip dary is equally divided into two components between
faults known to extend only to intermediate depths frontal accretion in a σ1 regime perpendicular to the
before branching on top of a deep crustal indenter margin and strike-slip motion parallel to the margin. The
(Laubscher, 1988; Daignières et al., 1989; Choukroune et same balance between a well-developed fold and thrust
al., 1989). In the same way, the El Pilar fault is here belt and a large parallel strike-slip fault is known to occur
assumed to overlie a high density mid-crustal wedge in California along the San Andreas fault (Mocent and
that acted as a backstop during migration of the defor- Suppe, 1987). Both the stress regime (σ1) and the deep
mation toward the south. Compressive deformation at geometry of this system are now well studied
the front of the wedge (Serranía and Maturín fold and (Wentworth et al., 1984; Wentworth and Zobak, 1989),
thrust belt) is mainly southward-directed, whereas the providing excellent examples for comparison with the
motion at the rear of the backstop is either strike-slip (the eastern Venezuelan margin. Nonetheless, only future
El Pilar dextral fault and related faults) (Jamison, 1991) or deep seismic profiles will be able to establish the depth
extensional (offshore northward-facing normal faults). where the El Pilar fault branches out on top of the
backstop and fold and thrust system, probably not much
Strain Partitioning along the South deeper than 12 km, on the basis of the present shallow
American–Caribbean Plate Boundary seismicity (Figures 2, 7).

A major component of strain partitioning is thus Composition of the High-Density Wedge


defined along the El Pilar fault, where continental
subduction induces decomposition of the oblique motion In addition to the problem caused by the high density
vector into two components. At the surface, one can required to fit the gravity anomaly, the composition of
observe almost pure strike-slip motion north of the El the deep crustal indenter is difficult to constrain. In the
Pilar fault and almost pure north-south shortening south Pyrenees, where two continental lithospheric plates
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 677

interact (the Iberian plate to the south and the European • The El Furrial and Orocual trends were formed
plate to the north), it is known that the brittle European later, providing closer traps for the oil generated
upper mantle acts as a backstop that progressively north of the present Pirital high, at a time when the
wedges the ductile Iberian lower crust out, forcing the Cretaceous source rocks were still immature in
thrusts to propagate southward. In the Alps, the upper these external units.
mantle of the hanging wall (the Apulian or African plate) • More recently, the tectonic overburial of the
probably helps to decouple the middle crust of the external units (related to the Pirital basement-
subducted plate (the European plate), but the composi- involved structure) has induced the maturation of
tion of the indenter itself is probably constantly modified the Cretaceous source rocks south of the Pirital
by the stacking (underplating) of fragments of European high, allowing a very young but short migration of
lower crust which cannot be subducted. oil toward the frontal structures.
In eastern Venezuela, where no continental block is
known to occur north of the El Pilar fault, the roots of the
early island arc system that developed above the ancient
oceanic subduction zone (Aves–Lesser Antilles arcs) REFERENCES CITED
could have participated in the formation of the backstop.
Nevertheless, it cannot be ignored that fragments of the Beck, C., 1984, On the mechanism of tectonic transport in
South American lithosphere itself (either upper mantle or zones of oblique subduction: Tectonophysics, v. 93, p. 1–11.
thinned lower crust) may also have contributed to the Beck, C., 1986, Collision Caraïbe, dérive andine et évolution
construction of the indenter. Only a small portion of the géodynamique Mésozoïque-Cénozoïque des Carïbes:
thinned crust may have been subducted. Revue de Geologie Dynamique et Geographie Physique, v.
27, p. 136–182.
Biju-Duval, B., A. Mascle, L. Montadert, and J. Wannesson,
1978, Seismic investigations in the Colombia, Venezuela
CONCLUSIONS and Grenada basins, and on the Barbados ridge for future
Despite the fact that deep seismic reflection and refrac- IPOD drilling: Geologie en Mijbouw, v. 57, p. 105–116.
Biju-Duval, B., A. Mascle, H. Rosales, and G. Young, 1983,
tion profiling in the eastern Venezuelan Coast Ranges is Episutural Oligo-Miocene basins along the north
only in the planning stages, the present integrated Venezuelan margin: AAPG Memoir 32, p. 347–358.
geologic and geophysical interpretation helps to constrain Bonini, W. E., 1978, Anomalous crust in the eastern Venezuela
a geodynamic model of oblique convergence. Like the basin and the Bouguer gravity anomaly field of northern
North Pyrenean fault in the Pyrenees or the Insubric line Venezuela and the Caribbean borderland: Geologie en
in the Alps, the El Pilar fault is assumed to overlie a high- Mijnbouw, v. 57, p. 117–122.
density mid-crustal wedge that acted as a backstop Bouysse, P., 1988, Opening of the Grenada back-arc basin and
during migration of the deformation toward the south. A evolution of the Caribbean plate during the Mesozoic and
minimum of continental subduction occurred until the early Paleogene: Tectonophysics, v. 149, p. 121–143.
late Pleistocene and is still active in Trinidad. At present, Bouysse, P., and D. Westercamp, 1990, Subduction of Atlantic
aseismic ridges and late Cenozoic evolution of the Lesser
the composition of the crustal wedge remains conjectural; Antilles island arc: Tectonophysics, v. 175, p. 349–380.
it could be made up of either Caribbean or South Burke, K., 1988, Tectonic evolution of the Caribbean: Annual
American lower crust and/or upper mantle. Review Earth and Planetary Sciences, v. 16, p. 201–203.
After Late Jurassic rifting, the South American passive Burke, K., C. Cooper, J. F. Dewey, P. Mann, and J. L. Pindell,
margin started to be deformed in the west, where the 1984, Caribbean tectonics and relative plate motions: GSA
Tethyan oceanic crust first disappeared. While the Lesser Memoir 162, p. 31–63.
Antilles or Aves arcs and related oceanic subduction Case, J. E., 1982, Major basins along the continental margin of
zones migrated eastward, an oblique convergent regime northern South America, in K. Burke and C. L. Drake, eds.,
was progressively initiated along the margin. This The geology of continental margins: New York, Springer
oblique convergence was the site of intense strain parti- Verlag, p. 733–742.
Chevalier, Y., J. F. Stephan, R. Blanchet, M. Gravelle, and A.
tioning, with coeval transpression and thrusting in the Bellizzia, 1985, L’ile de Margarita et la péninsule d’Araya:
south, and strike-slip or transtension to the north of the El (Venezuela); jalons d’une collision Crétacé supérieur sur la
Pilar fault, a near-surface expression of the South paléofrontière Caraïbes (Amérique du Sud) (abs.):
American–Caribbean plate boundary. Symposium Geologie des Caraïbes, Paris, Editions
Distinct episodes of petroleum generation and Technip.
migration took place during this complex geodynamic Chevalier, Y., J. F. Stephan, J. R. Darboux, M. Gravelle, H.
evolution of the Eastern Venezuelan basin (see Gallango Bellon, A. Bellizzia, and R. Blanchet, 1988, Obduction et
and Parnaud, 1995, for details): collision pre-Tertiaire dans les zones internes de la chaîne
Caraïbe vénézuélienne sur le transect de Margarita.
• In the early stages of the flexural evolution (prior to Péninsule d’Araya: Compte Rendu de l’ Academie des
Sciences, Paris, t. 307, série II, p. 1925–1932.
the formation of the frontal structures such as the Choukroune, P., et al., 1989, The ECORS Pyrenean deep
El Furrial or Orocual trends), early generated oils seismic profile; reflection data and the overall structure of
probably migrated over long distances from the an orogenic belt: Tectonics, v. 8, p. 23–39.
inner part of the belt toward the Orinoco to fill the Creole Petroleum Corporation, 1965, Geological maps D10
traps of the Faja Petrolífera. and D11, scale 1/100,000 scale.
678 Passalacqua et al.

Daignièåres, M., B. de Cabissole, J. Gallart, A. Hirn, E. Malavielle, J., and F. Cobb, 1986, Cinématique des déforma-
Surinach, and M. Torné, 1989, Geophysical constraints on tions ductiles dans trois massifs métamorphiques de
the deep structure along the ECORS Pyrenées line: l’Ouest des Etats-Unis: Albion (Idaho), Raft River et
Tectonics, v. 8, p. 1501–1558. Grouse Creek (Utah): Bulletín de la Societé Géologique de
Dewey, J. F., and J. L. Pindell, 1985, Neogene block tectonics of France, v. 8, p. 885–898.
eastern Turkey and northern South America: continental Maresch, W., 1974, Implication of a Mesozoic to early Tertiary
applications of the finite difference method: Tectonics, v. 4, collision-type plate-tectonic model in northern Venezuela
p. 71–83. for the southern Caribbean region: 7th Caribbean Geologic
Erlich, R. N., and S. F. Barrett, 1990, Cenozoic plate tectonic Conference, Guadeloupe, p. 485–491.
history of the northern Venezuela-Trinidad area: Tectonics, Martín, N., and E. Espinoza, 1990, Integración de la informa-
v. 9, p. 161–184. ción gravimétrica del flanco noreste de la cuenca oriental
Eva, A., K. Burke, P. Mann, P., and G. Wadge, 1989, Four- de Venezuela: V Venezuelan Geophysical Congress,
phase tectonostratigraphic development of the southern Caracas, p. 78–85.
Caribbean: Marine and Petroleum Geology, v. 6, p. 9–21. Mattauer, M., and B. Collot, 1986, Continental subduction,
Fernández, F., and H. Passalacqua, 1990, Procesamiento e thrusting and strike-slip faulting in the Canadian
interpretación de datos gravimétricos y magnéticos en la Cordillera: Bulletín de la Société Géologique de France, v.
cuenca oriental de Venezuela: V Venezuelan Geophysical 8, II, p. 899–910.
Congress, Caracas, p. 86–93. Moretti, I., and M. Larrére, 1989, LOCACE, computer aided
Fernández, F., and O. Pérez, 1974, Sismotectónica del área del construction of balanced cross-sections: Geobyte, v. 4,
Caribe y Occidente de Venezuela: Tesis de Grado, Univer- p. 16–24.
sidad Central de Venezuela, Facultad de Ingeniería, Mocent, V. S., and J. Suppe, 1987, State of stress near the San
Escuela de Geología y Minas, Caracas. Andreas fault: implications for wrench tectonics: Geology,
Fitch, T. J., 1970, Earthquake mechanisms in the Himalayan, v. 15, p. 1143–1146.
Burmese, and Andamam regions and continental tectonics Munro, S. E., and F. D. Smith, Jr., 1984, The Urica fault zone,
in central Asia: Journal of Geophysical Research, v. 75, northeastern Venezuela: GSA Memoir 162, p. 213–215.
p. 2699–2709. Orihuela, N., and R. Franklin, 1990, Modelaje gravimétrico de
Fourcade, E., J. Azema, F. Cecca, M. Bonneau, B. Peybernés, un perfil comprendido entre los poblados de Altagracia de
and J. Dercourt, 1991, Essai de reconstitution Orituco, Edo. Guarico y Caraballeda, Dtto. Federal,
cartographique de la paléogéographie et des paléo- envi- Venezuela: V Venezuelan Geophysical Congress, Caracas,
ronements de la Téthys au Tithonique supérieur (138 à 135 p. 466–472.
Ma): Bulletin de la Societé Géologique de France, v. 162, Passalacqua, H., A. S. Orange, J. Copley, C. Marquez, and R.
p. 1197–1208. B. Furgerson, 1990, Levantamientos magnetotelúricos
Gallango, O., and F. Parnaud, 1995, Two-dimensional experimentales en Venezuela para la exploración de hidro-
computer modeling of oil generation and migration in a carburos: V Venezuelan Geophysical Congress, Caracas,
transect of the Eastern Venezuelan basin, in A. J. Tankard, p. 118–125.
R. Suarez, and H. J. Welsink, Petroleum basins of South Pérez, O. J., and Y. P. Aggarwal, 1981, Present-day tectonics of
America: AAPG Memoir 62, this volume. the southeastern Caribbean and northeastern Venezuela:
Giraldo, C., 1990, Determinación del campo actual de Journal of Geophysical Research, v. 86, p. 10791–10804.
esfuerzos en la región Caribe a partir de datos Persool, K., 1990, Stratigraphic chart and tectonostratigraphic
sismológicos y neotectónicos: V Venezuelan Geophysical history, Trinidad area: K. Persad and Associates Ltd.,
Congres, Caracas, p. 70–77. Trinidad.
Jamison, W. R., 1991, Kinematics of compressional fold devel- Pimentel de Bellizzia, N., 1984, Mapa geológico estructural de
opment in convergent wrench terranes: Tectonophysics, v. Venezuela: Ministerio de Energía y Minas.
190, p. 209–232. Pindell, J. L., and S. F. Barrett, 1986, Geologic evolution of the
Kafka, A. L., and D. J. Weidner, 1981, Earthquake focal mech- Caribbean Regions; a plate tectonic perspective, in The
anisms and tectonic processes along the southern geology of North America, Vol. H, The Caribbean region:
boundary of the Caribbean plate: Journal of Geophysical GSA, p. 405–432.
Research, v. 86B, p. 2877–2888. Potié, G., 1989, La Serranía del Interior oriental sur le transect
Kellog , J. N., and W. E. Bonini, 1982, Subduction of the Cumaná-Urica et le bassin de Maturin (Venezuela): Ph.D.
Caribbean plate and basement uplifts in the overriding dissertation, University of Brest, Brest, France, 240 p.
with American plate: Tectonics, v. 1, p. 251–276. Rey, D., T. Quarta, P. Mouge, M. Miletto, R. Lanza, A.
Kugler, H. G., 1953, Jurassic to Recent sedimentary environ- Galdeano, M. T. Carrozzo, R. Bayer, and E. Armando,
ments in Trinidad: Bulletin der Vercinigung Schweiz- 1990, Gravity and aeromagnetic maps of the western Alps:
erischer Petroleum Geologen und Ingenieure, v. 20, contribution to the knowledge of the deep structures along
p. 27–60. the ECORS-CROP seismic profile, in F. Roure, R. F.
Ladd, J., 1976, Relative motion of South America with respect Heitzmann, and R. Polino, eds., Deep structure of the Alps:
to North America and Caribbean tectonics: GSA Bulletin, Mémoire de la Societé Geologique de France, París, 156;
v. 87, no. 7, p. 969–976. Mémoire de la Societé Géologique Suisse, Zurich, 1;
Laubscher, J., 1988, Material balance in Alpine orogeny: GSA volumen speciale della Societá Geologica Italiana, Roma, 1,
Bulletín, v. 100, p. 1313–1328. p. 107–121.
Lemiski, P. J., and L. D. Brown, 1988, Variable crustal Robertson, P., and K. Burke, 1989, Evolution of southern
structure of strike-slip fault zones as observed on deep Caribbean plate boundary, vicinity of Trinidad and
seismic reflection profiles: GSA Bulletín, v. 100, p. 665–676. Tobago: AAPG Bulletin, v. 73, p. 490–509.
Lillie, R. J., 1991, Evolution of gravity anomalies across colli- Rosales, H., 1972, La falla de San Francisco en el Oriente de
sional mountain belts: clues to the amount of continental Venezuela: IV Congreso Geológico Venezolano, Caracas
convergence and underthrusting: Tectonics, v. 10, 1969, Memorias Boletín de Geología, Caracas, v. 5,
p. 672–687. p. 2322–2336.
Crustal Architecture and Strain Partitioning in the Eastern Venezuelan Ranges 679

Roure, F., P. Choukroune, X. Berastegui, J. A. Muñoz, A. Wadge, G., and K. Burke, 1983, Neogene Caribbean plate
Villen, P. Matehron, M. Bareyt, M. Seguret, P. Camara, and rotation and associated Central American tectonic
P. Deramond, 1989, ECORS deep seismic data and evolution: Tectonics, v. 2, p. 633–643.
balanced cross-sections: geometric constraints on the Wadge, G., and D. Hudson, 1986, Neotectonics of southern
evolution of the Pyrenées: Tectonics, v. 8, p. 41–50. Tobago, in K. Rodrigues, ed., Transactions, First Geologic
Roure, F., R. Polino, and R. Nicholich, 1990, Early Neogene Conference of the Geologic Society of Trinidad and
deformation beneath the Po Plain: constraints on post-colli- Tobago, San Juan, Trinidad, p. 7–20.
sional Alpine evolution, in F. Roure, R. F. Heitzmann, and Wadge, G., and R. MacDonald, 1985, Cretaceous tholeites of
R. Polino, eds., Deep structure of the Alps: Mémoire de la northern continental margin of south America: the Sans
Societé Géologique de France, Paris, 156; Mémoire de la Souci Formation of Trinidad: Journal of the Geologic
Societé Géologique de Suisse, Zurich, 1; Volume speciale Society of London, v. 142, p. 297–308.
della Societá Geologica Italiana, Roma, 1, p. 309–322. Wadge, G., and J. B. Shepherd, 1984, Segmentation of the
Roure, F., J. O. Carnevali, Y. Gou, and T. Subieta, 1994, Lesser Antilles subduction zone: Earth and Planetary
Geometry and kinematics of the North Monagas thrust Sciences Letters, v. 71, p. 297–304.
belt (Venezuela): Marine and Petroleum Geology, v. 11, Watts, A. B., and M. Talwani, 1974, Gravity anomalies
no. 3, p. 347–362. seaward of deep-sea trenches and their tectonic implica-
Salvador, A., and R. M. Stainforth, 1968, Clues in Venezuela to tions: Geophysical Journal of the Royal Astronomical
the geology of Trinidad and vice versa: Transactions, IV Society, v. 36, p. 57–90.
Caribbean Geologic Conference, 1965, Trinidad, p. 31–40. Weeks, L. A., R. K. Lattimore, R. N. Harrison, B. G. Bassinger.,
Schubert, C., 1979, El Pilar fault zone northeastern Venezuela: and G. F. Merril, 1971, Structural relations among Lesser
brief review: Tectonophysics, v. 52, p. 447–455. Antilles, Venezuela and Trinidad-Tobago: AAPG Bulletin,
Schubert, C., 1981, Are the Venezuelan fault systems part of v. 55, p. 1741–1752.
the southern Caribbean plate boundary?: Geologische Wentworth, C. M., M. C. Blake, Jr., D. L. Jones, A. W. Walter,
Rundschau, v. 70, p. 542–551. and M. D. Zoback, 1984, Tectonic wedging associated with
Schubert, C., 1982, Origin of Cariaco basin, southern emplacement of the Franciscan assemblage, California
Caribbean sea: Marine Geology, v. 47, p. 345–360. Coast Ranges, in M. C. Blake, Jr., ed., Franciscan geology of
Schubert, C., 1984, Basin formation along the northern California: SEPM Special Publication 43,
Boconó–Morón–El Pilar fault system, Venezuela: Journal p. 163–173.
of Geophysical Research, v. 89, p. 5711–5718. Wentworth, C. M., and M. D. Zoback, 1989, The style of late
Soula, J. C., 1984: Genése de bassins sédimentaires en régime Cenozoic deformation at the eastern front of the California
cisaillant transcurrent: modéles expérimentaux et Coast Ranges: Tectonics, v. 8, p. 237–246.
exemples géologiques: Bulletin de la Société Geologique de Westbrook, G. K., and W. R. McCann, 1986, Subduction of
Belgique, v. 93, p. 83–104. Atlantic lithosphere beneath the Caribbean, in The geology
Speed, R. C., 1985, Cenozoic collision of the Lesser Antilles arc of North America, Vol. M, The western North Atlantic
and continental south America and the origin of the El region: GSA, p. 341–350.
Pilar fault: Tectonics, v. 4, p. 41–69. Wilson, C. C., 1968, The Los Bajos fault: IV Caribbean
Stein, S., J. Engeln, D. Wieuns, K. Fujita, and R. Speed, 1982, Geologic Conference, 1965, Trinidad, p. 87–89.
Subduction seismicity in the Lesser Antilles arc: Journal of Wong, J., H. Passalacqua, V. Graterol, and J. Flores, 1982,
Geophysical Research, v. 87, p. 8612–8664. Interpretación gravimétrica de la región Guanipa sur: I
Stephan, J. F., R. Blanchet, and B. Mercier de Lepinay, 1986, Congreso Venezolano de Geofísica, Caracas, p. 21–22.
Northern and southern Caribbean festoons (Panamá, Wu, S., A. E. Bally, and C. Cramez, 1990, Allochthonous salt,
Colombia-Venezuela and Hispaniola–Puerto Rico) inter- structure and stratigraphy of the north-eastern Gulf of
preted as pseudo-subductions induced by the eastwest Mexico, part II: structure: Marine and Petroleum Geology,
shortening of the peri-Caribbean continental frame, in F. C. v. 7, p. 334–370.
Wezel, ed., The origin of arcs, Developments in Geotec-
tonics: Amsterdam, Elsevier, no. 21, p. 401–422. Authors’ Mailing Addresses
Sykes, L. R., W. R. McCann, and A. L. Kafka, 1982, Motion of H. Passalacqua
Caribbean plate during last 7 millon years and implica-
F. Fernandez
tions for earlier Cenozoic movements: Journal of Geophys-
ical Research, v. 87 p. 10656–106676. Intevep, S.A.
Sylvester, A. G., and R. R. Smith, 1976, Tectonic transpression Apartado postal 76343
and basement controlled deformation in San Andreas fault Caracas 1070-A
zone, Salton trough, California: AAPG Bulletin, v. 60, Venezuela
p. 2081–2102.
Van den Driessche, J., 1986, Cinématique de la déformation Y. Gou
ductile dans la cordillére canadienne: relations chevauche- Beicip
ments-décrochements: Bulletin de la Societé Geologique de B.P. 213, 92502
France, v. 8, p. 911–920. Rueil-Malmaison
Vierbuchen, R. C., 1984, The geology of the El Pilar fault zone
France
and adjacent areas in northeastern Venezuela, in W. E.
Bonini,. R. B. Horgroves, and R. Shagam, eds., The south
Caribbean–South America plate boundary and regional F. Roure
tectonics: GSA Memoir 162, p. 189. IFP
Vignali, M., 1979, Estratigrafía y estructura de las cordilleras B.P. 311,92506
metamórficas de Venezuela Oriental (Península de Rueil-Malmaison
Araya–Paria y Isla de Margarita): GEOS, v. 25, p. 19–66. France
Stratigraphic Synthesis of Western Venezuela

François Parnaud Maria Angela Capello


Yves Gou Irene Truskowski
Jean-Claude Pascual Herminio Passalacqua
Beicip-Franlab Intevep, S.A.
Petroleum Consultants Caracas, Venezuela
Rueil-Malmaison, France

Abstract

T he sedimentary basins of western Venezuela contain large volumes of oil. However, most of the large
structures have already been produced. Exploration for new reserves of light and medium oil now
depends on integrated studies that lead to a more comprehensive basin evaluation. This paper presents an
integrated account of the Lake Maracaibo and Barinas-Apure basins of western Venezuela. It is a fully inte-
grated study but focuses on the genetic and seismic stratigraphy of more than 600 wells, reference outcrops,
and 4000 km of reflection seismic data.
Six unconformity-bounded supersequences record the dynamics of Mesozoic–Cenozoic basin evolution
from extension to collision. Supersequence A was deposited during an episode of Jurassic rifting, and superse-
quence B corresponds to the subsequent Early–Late Cretaceous passive margin. Supersequence C marks the
transition to a compressive regime in the Late Cretaceous and early Paleocene. Compression resulted from
collision and obduction of the Pacific volcanic arc with the South American plate. Supersequence D records the
development of the late Paleocene–middle Eocene foreland basin in front of the volcanic arc and emplacement
of the Lara nappes. Supersequences E and F are attributed to modification of the foreland basin by late
Eocene–Pleistocene collision of the Panamá arc. The uplifted Serranía de Perijá, Macizo de Santander, and
Mérida Andes partitioned the foreland basin, creating the present Lake Maracaibo and Barinas-Apure basins.
Supersequence B contains the Cretaceous La Luna source rock (sequences K3, K4, K5). The Colón and
Burgüita formations form the principal supersequence C seals (sequence K6). The principal reservoir units
occur in supersequence D, including the prolific Eocene Misoa and Gobernador formations (sequences T1, T2).
Reservoirs of the La Rosa and Lagunillas formations occur in supersequence F and in the Betijoque Molasse.

Resumen

L as cuencas sedimentarias del Occidente de Venezuela contienen inmensas acumulaciones de hidrocar-


buros. Sin embargo, la mayoría de las estructuras grandes ya han sido explotadas. La exploración que
incorpore más reservas de crudo liviano y mediano debe entonces basarse actualmente en estudios integrados
que conduzcan a una evaluación más exhaustiva de estas cuencas. Este artículo presenta una vision integrada
de las cuencas del Lago de Maracaibo y Barinas-Apure, en el occidente de Venezuela. Es un estudio totalmente
integrado y enfocado especialmente a la estratigrafía genética y sísmica de más de 600 pozos, afloramientos de
referencia y alrededor de 4000 km de líneas sísmicas.
Seis supersecuencias, limitadas por discordancias, evidencian la dinámica evolución de las cuencas en el
Mesozoico–Cenozoico, de un proceso de extensión a uno de colisión. La supersecuencia A fue depositada
durante un episodio de apertura de corteza del Jurásico. La supersecuencia B corresponde al margen pasivo
subsiguiente, durante el Cretácico Temprano al Tardío. La supersecuencia C marca la transición a un régimen
compresivo en el Cretácico Tardío y Paleoceno Temprano. La compresión es el resultado de la colisión y
obducción del arco volcánico pacífico al Oeste con la placa suraméricana. La supersecuencia D pone de mani-
fiesto el desarrollo de la cuenca de antepaís del Paleoceno Tardío–Eoceno Medio, al frente del arco volcánico
pacífico, y el emplazamiento de las napas de Lara. Las supersecuencias E y F se atribuyen a las modificaciones
en la cuenca de antepaís debidas a la colisión Eoceno Tardío–Pleistoceno del arco de Panamá. Los levan-
tamientos de la Serranía de Perijá, del Macizo de Santander y de los Andes de Mérida particionaron la cuenca
de antepaís generando así las actuales cuencas del Lago de Maracaibo y Barinas-Apure.
La supersecuencia B contiene la roca madre La Luna de edad Cretácico. Las formaciones Colón y Burgüita
coforman los sellos principales de la supersecuencia C. Las principales unidades reservorio se ubican en la
supersecuencia D, incluyendo a las prolíficas formaciones Misoa y Gobernador del Eoceno. Las formaciones La
Rosa y Lagunillas generan reservorios dentro de la supersecuencia F en la cual se ubica la sedimentación
molássica de la formación Betijoque.

Parnaud, F., Y. Gou, J.-C. Pascual, M. A. Capello, I. Truskowski, and H. Passalacqua, 681
1995, Stratigraphic synthesis of western Venezuela, in A. J. Tankard, R. Suárez S., and
H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 681–698.
682 Parnaud et al.

INTRODUCTION
Western Venezuela can be divided into several struc-
tural units (Zambrano et al., 1970) (Figure 1): (1) the
Guyana shield or Cuchivero granitic province
(Menéndez, 1968); (2) the Mérida Andes mountain range,
which separates the Barinas-Apure basin to the southeast
from the Lake Maracaibo basin in the northwest; (3) the
Serrania de Perijá in the west, which separates the Lake
Maracaibo basin (east) from the Colombian César-
Ranchería basin (west); and (4) the Serrania de Trujillo,
which separates the Lake Maracaibo basin from the Lara
nappes (Stephan, 1985).
These structural units record a long history of basin
evolution (Figure 2). The earliest episode involved the
progressive evolution from extension to a Caribbean–
Tethyan passive margin at the edge of the South
American plate. This evolution spanned late Triassic–
Cretaceous time. It was followed by collision of the
Pacific plate with the South American plate and
building of a Late Cretaceous-Paleocene mountain range
with an associated foreland basin. Collision and
migration of the Caribbean plate since the Paleocene
resulted in the Lara thrust belt and the Eocene foreland
basin. The subsequent Andean orogeny is attributed to
collision of the Panamá arc. This collision also parti-
tioned the Cretaceous passive margin into the post-
middle Miocene Lake Maracaibo and Barinas-Apure
basins.
This geologic history is expressed in a hierarchy of Figure 1—Location of the study area and stratigraphic
depositional sequences. On a large scale, the stratigraphy syntheses of Barinas-Apure and Lake Maracaibo basins.
can be divided into Paleozoic and Mesozoic–Cenozoic Cross sections: A–A' is in Figure 3, B–B' and C–C' in
successions. Jurassic extension records separation of Figure 5, and I–I' and J–J' in Figure 15. Circled numbers
North and South America (Pindell and Erikson, 1993; refer to the locations of these numbered figures.
Parnaud et al., 1994). The Mesozoic–Cenozoic succession
contains a suite of unconformity-bounded sequences that
describe the dynamics of basin evolution along the structural–tectonic modifications (Figure 3). Sedimento-
northern part of the South American plate (Figures 2, 3), logically, the Cretaceous and Paleocene consist of a
as follows: heterogeneous carbonate-siliciclastic association. In
contrast, the Cenozoic is represented mainly by a silici-
• Supersequence A resulted from an episode of clastic sedimentary system.
Jurassic rifting. Dating of the sequence boundaries and the maximum
• Supersequence B corresponds to the Early–Late flooding surfaces was established with available strati-
Cretaceous stage of passive margin development. graphic and biostratigraphic information and by
• Supersequence C was a transitional phase of the comparison with the global sea level charts of Haq et al.
Late Cretaceous–early Paleocene passive margin (1987). We believe that tectonism due to emplacement of
deposition behind the compressive arc. the Lara nappes modified the eustatic signature, particu-
• Supersequence D was deposited in a foreland basin larly in the flexured zones such as those located in front
during the late Paleocene–middle Eocene, when of the deformation and along the zone of lateral ramps
collision and obduction of the Pacific volcanic arc that limit the nappes’ extent along the eastern border of
overrode the South American plate and emplaced the lake.
the Lara nappes.
• Supersequences E and F were related to the late
Eocene–Pleistocene phase of foreland basin subsi-
dence caused by collision of the Panamá arc. This
PALEOZOIC SUCCESSION
episode of deformation was responsible for separa- The sequences deposited during the Paleozoic were
tion of the Lake Maracaibo and Barinas-Apure identified in several areas, in particular, the Guyana
basins. shield, Mérida Andes, Lake Maracaibo basin, and
Serrania de Perijá. A stratigraphic column for the
Internally, these six supersequences comprise a series Paleozoic of western Venezuela has been established
of minor sequences that reflect eustatic processes and (González de Juana, 1980).
Stratigraphic Synthesis of Western Venezuela

Figure 2—Geotectonic evolution of Venezuelan supersequences A–F. Symbols: M, Machiques; U, Uribante; T, Trujillo; L, Lake Maracaibo; CC,
Central Cordillera; G, Guajira; P, Paraguana; Ti, Trinidad. (Modified after Pindell and Erikson, 1993.)
683
Figure 3—Stratigraphic chart (see Figure 1, A–A', for location). Legend: 1, continental sandstones and shales; 2, inner to middle shelf carbonates; 3, middle to
outer shelf carbonates and shales; 4, outer shelf to bathyal shales; 5, nearshore regressive sandstones; 6, nearshore transgressive sandstones; 7, phtanites.
Stratigraphic Synthesis of Western Venezuela 685

The lower Paleozoic in the southern flank of the Mérida


Andes consists of fossiliferous siltstones of the Ordovician
Caparo Formation (Christ, 1927) and the Silurian El
Horno Formation (Martín Bellizia, 1968). The middle
Paleozoic Río Cachirí Group of Liddle (1928) occurs in the
Serrania de Perijá. This group consists of marine shelf
sedimentary rocks that are rich in fauna, such as
brachiopods and pelecypods. The upper Paleozoic in the
Mérida Andes is stratigraphically diverse. The Mucu-
chachí Formation (Christ, 1927) records marine inunda-
tions, the Sabaneta Formation (Oppenheim, 1937) shows
an episode of continental deposition, and the Palmarito
Formation (Christ, 1927) is evidence of a final marine
event. In the Serrania de Perijá, the upper Paleozoic is
represented by two sets. A lower set includes the Caño
del Noroeste, Caño Indio, and Río Palmar formations
which are believed to be equivalent to the Sabaneta
Formation toward the east in the Mérida Andes. A
younger set, containing the Palmarito Formation, is
composed of sandy facies and marine limestones of inner
shelf origin. In the western Venezuelan basins, no well
has thus far reached this stratigraphic level.
Nevertheless, in the subsurface of the Barinas-Apure
basin south of Apure (Elorza zone), a sequence has been
interpreted in seismic sections with a thickness of 2 sec
(about 4500 m). This sequence is characterized by parallel
reflections that are relatively continuous beneath the
Cretaceous section (Figure 4A). Its top corresponds to a
regional unconformity that is well defined by trunca-
tions, while its base is difficult to recognize. A Paleozoic
age is inferred on the basis of regional correlations and
the structural style that affects it. This Paleozoic succes-
sion was subjected to a strong compressive event that
resulted in thrusts and fault-bend folds. These structures
were subsequently peneplaned and unconformably
covered by Mesozoic strata (Figure 4A). The structural
style distinguishes the Paleozoic rocks from the Jurassic
sequences.
In the Lake Maracaibo basin, drilling encountered
metamorphic rocks beneath the Cretaceous. These strati-
graphic levels were considered by González de Juana et
al. (1980) to be possible equivalents of Paleozoic forma-
tions in the central Andean region. Nevertheless, seismic
data do not record a Paleozoic stratigraphy in the Lake Figure 4—Seismic profiles. (A) Paleozoic succession. (B)
Maracaibo basin similar to that interpreted for the Triassic–Jurassic extensional supersequence A. See
Barinas-Apure basin. We believe that this may reflect the Figure 1 for locations.
level of metamorphism.
The hydrocarbon potential of the Paleozoic succession
is poorly known. Alberdi et al. (1994) showed that the Jurassic Supersequence A: Extension
organic matter of the Palmarito Formation samples has a Supersequence A occurs in outcrops of Perijá and
high maturity level, which precludes an accurate evalua- Mérida Andes and in the subsurface of the Lake
tion of its real initial hydrocarbon potential. Maracaibo and Barinas-Apure basins. In Perijá, superse-
quence A forms the La Ge Group (Hea and Whitman,
1960) and includes the Tinacoa (Liddle et al., 1943),
Macoíta (Hedberg and Sass, 1937), and La Quinta
MESOZOIC–CENOZOIC SUCCESSION (Künding, 1938) formations. These formations reflect sedi-
The Mesozoic–Cenozoic succession results from the mentation in continental environments that were locally
Jurassic rift phase attributed to the fragmentation of sourced by volcanic material such as volcanic ashes. In the
Pangea and from the Cretaceous–Tertiary phase of Mérida Andes, this Jurassic supersequence is identified as
collision between Pacific and South American plates the La Quinta Formation, which was also deposited in a
(Figure 2). continental setting with conspicuous red sandstones.
686 Parnaud et al.

In the subsurface of the Barinas-Apure basin, it is Táchira, and the Barquisimeto trough in Trujillo. In
difficult to identify supersequence A in seismic sections addition to these troughs, this sequence forms a wide-
without well control. However, southwest of the Barinas spread cover, except in the southwestern part of Apure.
basin, where outcrops show a significant development of Hedberg (1931) initially described sequence K0 in the
this Triassic–Jurassic sequence, some seismic lines Negro River area (Serrania de Perijá), and it was later
display a 1-sec-thick (3000-m) sequence beneath the named the Río Negro Formation (Hedberg and Sass,
Cretaceous (Figure 4B). The base of the supersequence is 1937). It marks the basal continental component of the
obscure, but its top is expressed as a strong unconform- Cretaceous passive margin basin. Although it is wide-
ity. Its seismic signature consists of discontinuous reflec- spread, its age is poorly constrained; available evidence
tions of variable amplitudes. The structural style of these suggests a Neocomian–Aptian age. Seismic resolution is
intervals is generally characterized by normal faults that insufficient to establish the upper bounding surface with
bound small half-grabens. confidence.
In the subsurface of the western part of Lake This Lower Cretaceous sequence has a low hydro-
Maracaibo basin, west of the Icotea structural trend, a carbon potential. Source rocks are unknown, and
half-graben is observed in seismic sections and has an although it has a high sandstone content, reservoir char-
eastward-dpping basin-forming fault. The base of super- acteristics are generally believed to be poor.
sequence A is not recognized, but its upper surface is
marked by truncated reflections. Wells indicate that this Aptian Sequence K1
sequence corresponds to the La Quinta Formation. Continental Rio Negro deposition was terminated by
Because of an apparent absence of source rocks and a Cenomanian–Campanian marine transgression that
poor reservoir characteristics, supersequence A has little flooded the Guyana cratonic platform (Figure 2). This
hydrocarbon potential. There is no production from transgression was episodic as evidenced by a back-
Jurassic levels in the study area. stepping suite of depositional sequences, the first of
which has an Aptian age (Figure 3). The Aptian Apón
Cretaceous Supersequence B: Passive sequence (K1) (Sutton, 1946) is characterized by shallow
Margin marine shelf sedimentation and displays lateral facies
changes. Toward the east (Mérida Andes), littoral sand-
At the beginning of the Cretaceous, a marine trans-
stones form the basal part of the Peñas Altas Formation
gression caused inundation of the Guyana shield. This
(Renz, 1959) (Figure 6). This sequence consists of three
transgression is correlated to the eustatic changes that
parts (Figure 3):
occurred worldwide and lasted until Cenomanian–
Campanian time (Figure 2). Sporadic volcanic material • The lower part is interpreted as a transgressive
within the La Luna Formation suggests the presence of a systems tract (TST) and comprises several retro-
volcanic arc toward the west, implying subduction of the gradational parasequences. This TST corresponds
Pacific plate. The apparent reduction of fault-controlled to the Tibú Member and was deposited in an inner
subsidence, the overall transgressive deepening of the shelf environment where littoral bioclastic bars
basin, and the stratigraphy suggest that Cretaceous developed.
supersequence B was deposited as a passive margin • The middle part forms the maximum flooding
terrace wedge behind a volcanic arc. The passive margin surface (MFS) and include the Machiques Member
phase ended with collision of the Pacific arc and the and its laterally equivalent, the Guáimaros
South American plate and flexural subsidence of Member. Both were deposited in middle shelf envi-
foreland basins. ronments with several intercalations of shallower
During this period of passive margin development, deposits. Based on previous work, an interval rich
several sequences were deposited (Figure 3). Together in Orbitolina texana is equated with the middle
they formed the terrace wedge of supersequence B. The Aptian event (the 111-Ma MFS of Haq et al., 1987).
structural style is characterized by steeply dipping • The upper part of this sequence is characterized by
reverse faults of post-Cretaceous age. Structural a highstand and progradation. This regressive part
inversion of older normal faults are interpreted from is the Piché Member of the Lake Maracaibo zone,
seismic data (Figure 5, C–C’). which was deposited in an inner shelf environ-
In general, it is difficult to differentiate these passive ment. In the Mérida Andes, the central part of the
margin sequences seismically, except for local reflection Peñas Altas Formation was deposited in a littoral
truncations (e.g., onlaps and downlaps) in an otherwise setting in which bars developed.
continuous supersequence B with persistent subparallel
reflections. The internal seismic sequence boundaries In the Lake Maracaibo succession, reflection seismic
separate six depositional sequences, K0 through K5, as data does not clearly distinguish this sequence from
discussed below. other Cretaceous units. Nevertheless, its base does show
local onlaps and its top shows weak truncations. The
Neocomian–Barremian Sequence K0 internal geometry of sequence K1 is marked by three
In the Early Cretaceous, a thick sequence of conti- high-amplitude reflectors (Figure 7A). The thickness of
nental sediments was deposited in three troughs: the this sequence varies up to 300 m, but it is absent in the
Machiques trough in Perijá, the Uribante trough in Barinas-Apure basin (Figures 5, B–B’, and 6). The
Stratigraphic Synthesis of Western Venezuela

Figure 5—(Top) Tectonostratigraphic cross section B–B' of Barinas-Apure basin. (Bottom) Tectonostratigraphic cross section C–C' of Lake
Maracaibo basin. See Figure 1 for locations.
687
688 Parnaud et al.

Figure 6—Paleogeography of the Aptian depositional


sequence K1. Legend: 1, Lara nappes, actual position; 2,
positive areas; 3, nearshore clastics; 4, inner shelf carbon-
ates and shales; 5, middle shelf carbonates and shales; 6,
thickness contours in feet.

estimated sedimentation rate is about 25 m/m.y.


The Aptian depositional sequence has good hydro- Figure 7—Seismic profiles. (A) Aptian depositional
carbon potential in the western part of the Lake sequence K1 showing onlaps and truncations. (B) Eocene
Maracaibo basin. The Machiques source rock is rich in supersequence D showing top and base bounding uncon-
type II organic matter (Alberdi et al., 1994). The reservoir formities. Columns: 1, depositional sequences; 2, superse-
intervals are believed to have dissolution and fracture quences. See Figure 1 for locations.
porosity.

Albian–Lower Cenomanian Sequence K2 Formation were deposited in an inner shelf setting


with fringing shoreline facies (Figure 8).
The second major marine transgression took place • The middle part corresponds to the maximum
during the Albian, invading the entire study area from flooding surface (the 97-Ma MFS of Haq et al.,
the Serrania de Perijá to the southeastern limit of the 1987) developed in the S Member of the Escan-
Barinas-Apure basin and toward the Guyana shield dalosa Formation in the Barinas-Apure basin. In
(Figure 2). This depositional sequence (K2) includes the the Lake Maracaibo basin, this MFS is absent due to
Lisure (Rod and Maync, 1954), Maraca (Rod and Maync, erosion (Figure 9). There the shales of the S
1954), La Puya (Renz, 1959), lower Capacho (Sievers, Member were deposited in an open marine middle
1888), Aguardiente (Notestein et al., 1944), and basal shelf environment.
Escandalosa (Renz, 1959, S and R members) formations. • The upper part of the sequence is a progradational
This sequence has a threefold subdivision (Figure 3): highstand systems tract. It is related to the regres-
sive sandstones of the R Member in the Escan-
• The lower part is a transgressive systems tract that dalosa Formation and was deposited in a shallow
was internally built by retrograde parasequences. littoral environment with shoreline, barrier bar,
During this TST, the Lisure Formation was and coastal lagoon sedimentation. The lateral
deposited in a middle shelf environment. At the equivalent in the Lake Maracaibo area is absent
same time, transgressive sands of the Aguardiente due to erosion.
Stratigraphic Synthesis of Western Venezuela 689

Figure 8—Paleogeography of the Albian–lower Ceno- Figure 9—Paleogeography at top of Albian–lower Ceno-
manian depositional sequence K2. Legend: 1, Lara nappes, manian depositional sequence K2. Legend: 1, Lara nappes,
actual position; 2, positive areas; 3, nearshore clastics; 4, actual position; 2, positive areas; 3, nearshore clastics; 4,
inner to middle shelf carbonates and shales; 5, thickness inner shelf sandstones and carbonates; 5, middle shelf
contours in feet. carbonates and shales.

In Perijá and Lake Maracaibo, Canache et al. (1994) data from the Lake Maracaibo area also shows basal
identified a hiatus of early Cenomanian age between onlaps (Figure 7A), as well as an irregular upper surface
sequence K2 and the overlying upper Cenomanian strata that we interpret as a paraconformity.
of sequence K3 (Figures 3, 9). Erosion has partially This sequence, which is regionally persistent (Figure
truncated the Albian sequence. We attibute this hiatus to 8), has a thickness varying up to 600 m. Estimated sedi-
collision of the Pacific volcanic arc with South American mentation rates are 50 m/m.y. in the southern part of the
continental crust and to flexural deformation in front of Mérida Andes and 12 m/m.y. in the shelf zone.
the applied load. A foreland basin is located west of The hydrocarbon potential of sequence K2 lies in the
Serrania de Perijá, and an associated forebulge occurs in Lisure reservoirs with their fracture porosity and in
the Perijá and Lake Maracaibo areas. Forebulge uplift sandstone reservoirs in the Aguardiente (TST) and
resulted in emergence, restricted deposition, and erosion Escandalosa (R Member, HST) formations that preserve
of the upper part of sequence K2 during the early Ceno- intergranular porosity.
manian. This is reflected in a regressive wedge in the
Mérida Andes and in deposition of the highstand regres- Upper Cenomanian–Lower Campanian
sive Escandalosa Sandstone. Following this event, a new Sequences K3, K4, and K5
trangression flooded the entire area. This transgression Foreland downwarping following deposition of Early
probably resulted from renewed compression and Cretaceous sequences resulted in episodic late Ceno-
regional flexural downwarping. Basinal accumulation of manian–early Campanian transgression and three back-
shaly calcareous facies of the La Luna Formation stepping depositional sequences: K3, K4, and K5 (Figure
initiated late Cenomanian sedimentation. 2). These sequences are present in the Serrania de Perijá
In Barinas seismic sections, identification of the and Lake Maracaibo basin as the La Luna Formation and
sequence K2 is generally tenuous. However, there are the Tres Esquinas Member (Stainforth, 1962), in the
some onlaps of the transgressive systems tract (Aguardi- Mérida Andes as the Capacho (upper Seboruco and
ente Formation) and local downlap of the highstand Guayacán members; Sievers, 1888) and La Luna forma-
systems tract (R Member of Escandalosa Formation) tions, and in the Barinas-Apure basin as the Escandalosa
associated with a gentle unconformity at its top. Seismic (P and O members; Renz, 1959) and Navay (La Morita and
690 Parnaud et al.

• There are two well-defined facies tracts (Figure 10).


In the west, the middle shelf–bathyal tract contains
calcareous and shaly facies of the La Luna
Formation. Eastward detrital littoral sediments of
the Escandalosa Formation (P member) and Navay
Formation (Quevedo Member) were deposited
(Figure 11).

These sequences were established seismically on the


basis of onlap relationships. However, no particular
internal geometric characteristics were observed (Figure
7A). The thickness of the sequences varies from 150 to
>600 m (Figure 10). Sedimentation rates of 30 m/m.y. in
the southern Andes and 8 m/m.y. in the Maracaibo shelf
area are estimated.
This succession has substantial hydrocarbon potential,
including the exceptional La Luna source rock (Alberdi
et al., 1994) and favorable reservoirs with intergranular
porosities in the Escandalosa (TST) and Navay (Quevedo
Member, HST) formations.

Late Cretaceous-Paleocene Supersequence C:


Passive to Active Margin Transition
In the Late Cretaceous, a new phase in tectonic
evolution was marked by collision of the Pacific volcanic
arc with the South American plate. This collision trans-
formed the passive margin into an active belt, creating a
Figure 10—Paleogeography of upper Cenomanian–lower foreland basin with an associated foredeep to the west
Campanian depositional sequences K3, K4, and K5. (Perijá area) and a forebulge in the Barinas area. Never-
Legend: 1, Lara nappes, actual position; 2, positive areas; theless, toward the north and northeast, the passive
3, nearshore clastics; 4, internal to middle shelf sand- margin setting persisted until emplacement of the Lara
stones and carbonates (Guayacan Member); 5, outer shelf thrust belt and nappes. This history indicates a scissor-
to upper bathyal carbonates and shales; 6, thickness
contours in feet.
type closure of the old passive margin through the Late
Cretaceous and early Paleocene (Figure 2). This transi-
tional phase was also characterized by a regression that
Quevedo members; Kehrer, 1938; Renz, 1959) formations. resulted in three depositional sequences, K6, K7, and K8,
Several maximum flooding surfaces are recognized. as discussed below (Figure 3).
The upper Cenomanian interval at the base of the La
Luna Formation is not obvious. Haq et al. (1987) docu- Upper Campanian–Maastrichtian Sequence K6
ment their 92-Ma MFS as a prominent event. In contrast,
the next MFS at the base of the La Morita Member is well Regression began at the beginning of the Late Creta-
defined in the Barinas-Apure basin (equivalent to the ceous. Simultaneously, toward the west, the Pacific
88-Ma datum of Haq et al., 1987). The third MFS occurs volcanic arc collision formed a foredeep within which
in the Tres Esquinas Member in the Lake Maracaibo shaly facies of the Colón Formation were deposited
basin (79-Ma datum of Haq et al., 1987). The highstand (Liddle, 1928). The associated forebulge migrated from
systems track of each sequence and the corresponding the Lake Maracaibo depocenter to Barinas-Apure, where
sequence boundaries are not precisely determined. arenaceous shoreline facies of the Burgüita Formation
The characteristics of sequences K3, K4, and K5 are as were deposited (Renz, 1959) (Figure 12). Sedimentation
follows: of this sequence (K6) ended in a highstand systems tract
that is expressed in the Mito Juan Formation (Garner,
• Volcanic ash beds at the base of the La Luna 1926). The lower and upper boundaries of this sequence
Formation suggest the presence of the Pacific are assigned to the late Campanian and late Maas-
volcanic arc west of the study area. trichtian (Figure 3).
• The basin deepens rapidly from inner shelf to Several smaller scale depositional units build the
bathyal depths, possibly reflecting eastward internal fabric of sequence K6, with indeterminate
migration of the forebulge toward the Barinas- flooding surfaces. Generally the shaly units of the Colón
Apure basin complex. Formation are interpreted as transgressive drapes and
• The three transgressive sequences each culminate the sandier stratigraphy of the Mito Juan Formation as
in maximum flooding surfaces. highstand progradational depositional systems.
Stratigraphic Synthesis of Western Venezuela 691

Figure 12—Paleogeography of upper Campanian–Maas-


trichtian depositional sequence K6. Legend: 1, Lara
nappes, actual position; 2, positive areas; 3, inner to
middle shelf clastics; 4, outer shelf shales and scarce
sandstones; 5, thickness contours in feet.

Internal clinoform geometries indicate local eastward


progradation of the Mito Juan Formation (Figure 13B).
Eroded topsets suggest latest Cretaceous tectonism in the
Perijá area. This depositional sequence is generally more
argillaceous than the previous stratigraphy and conse-
quently less competent. Because of this, the brittle fault
styles of the lower sequences differ from that in the
Campanian–Maastrichtian rocks, where disharmony
occurs. The thickness of the sequence varies up to 900 m
(Figure 12). Sedimentation rates of 65 m/m.y. in the shelf
and 150 m/m.y. in the Perijá foredeep are estimated.
Figure 11—Genetic stratigraphy of Cretaceous succession. Sandstones of the Burgüita Formation (TST) in Barinas
See Figure 1 for location. and the Mito Juan Formation (HST) offer reservoir
potential.
In the Barinas-Apure basin, this sequence is poorly Upper Maastrichtian–Lower Paleocene
distinguishable on seismic profiles (Figure 13A). It is thin Sequences K7 and K8
and has no obvious seismic signature apart from onlap
geometries and truncations beneath the upper bounding Toward the end of the Cretaceous, the Perijá foredeep
surface. In the central areas, the sequence is deeply was filled with highstand sediments of the Mito Juan
eroded and discontinuous. In the Lake Maracaibo basin, sequence (K6), which were sourced from the west. The
the base of sequence K6 corresponds to a continuous entire area was affected by erosion above shallowing
strong reflector (Figure 7B). However, its upper surface is basement. A new transgressive episode from the
less obvious, except toward the west, where it has trun- northeast deposited two subordinate Paleocene
cation relationships (Figure 13B). sequences, K7 and K8 (Figure 3). The lower sequence
692 Parnaud et al.

Figure 14—Paleogeography of the upper Maastrichtian–


lower Paleocene depositional sequences K7 and K8.
Legend: 1, Lara nappes, actual position; 2, positive areas;
3, continental to deltaic clastics; 4, inner to outer shelf
Figure 13—Seismic profiles. (A) Supersequences B and C carbonates and shales; 5, bathyal with turbidites, shales,
showing basement topography and Cretaceous onlaps. and scarce sandstones; 6, thickness contours in feet.
(B) Paleocene fans in the Marcelina Formation (deposi-
tional sequences K7 and K8). Columns: 1, depositional
sequences; 2, supersequences. See Figure 1 for location.
The base of the shelf depositional system in the
Maracaibo basin is characterized seismically by an
covered the entire shelf and displays marine characteris- erosional surface (Figure 13B). The succeeding deltaic
tics, while the upper sequence is essentially deltaic. system has discontinuous and strong amplitude reflec-
The shelf terrace wedge (K7) comprises several forma- tions; its base is marked by onlaps and downlaps and its
tions. The Guasare Formation (Garner, 1926) consists of top by local truncations. This deltaic sequence thickens
shallow marine deposits in the Lake Maracaibo basin. toward the west where it locally forms delta-front fans.
The Trujillo Formation (Hodson, 1926) comprises deeper These Paleocene rocks are up to 600 m thick (Figure 14).
marine deposits northeast of the lake area, while the An average sedimentation rate of 30–80 m/m.y. is
Catatumbo Formation (Notestein, 1944) was built by estimated.
deltaic sedimentation toward the south (Figure 14). The The hydrocarbon potential of these sequences is
overlying deltaic suite (K8) contains three formations: the restricted to the terrigenous clastics of the Barco
Barco and Los Cuervos formations to the south and the Formation (TST). The carbonaceous rocks of the
Marcelina Formation to the north. Toward the northeast, Marcelina Formation (HST) are unlikely to offer source
lowstand facies of the Trujillo Formation (Hodson, 1926) potential as expected (Alberdi et al., 1994).
developed.
We have mapped three sedimentary domains in each
sequence (Figure 14). The first is in the southwest and Upper Paleocene–Middle Eocene
consists of sandstones and mudstones of the Orocue Supersequence D: Collisional Basins
Group (Notestein, 1944) and Marcelina Formation. The
second occurs in the central area, where shallow marine Emplacement of the Lara nappes began north of the
bioclastic and calcareous sediments of the Guasare Lake Maracaibo basin at the end of the Paleocene (Figure
Formation were deposited. The third sedimentary 2). These nappes gradually encroached eastward,
domain, located toward the north, contains bathyal forming new foreland basins. One of these trends
sediments of the Trujillo Formation. N 20º W, parallel to the northeastern margin of Lake
Stratigraphic Synthesis of Western Venezuela 693

Figure 15—(Top) Cross section I–I' showing stratigraphic


relationships among Upper Cretaceous–lower Tertiary
sequences K7, K8, and T1. (Bottom) Cross section J–J'
showing stratigraphic relationships between middle
Eocene depositional sequences T2 and T3 in the Barinas-
Apure basin. See Figure 1 for locations.

Maracaibo; another is oriented approximately east-west


in front of the nappes. This flexural deformation is
reflected in a suite of transgressive and regressive cycles
of Eocene age. The base and top of supersequence D
correspond to regional unconformities that are expressed
seismically by numerous onlaps and truncations (Figure
7B). Intermittent subsidence and a possible eustatic
overprint resulted in three depositional sequences, T1,
T2, and T3, as discussed below (Figure 3).

Upper Paleocene–Lower Eocene Sequence T1


There are two parts to this upper Paleocene–lower
Figure 16—Genetic stratigraphy of the Misoa Formation.
Eocene Sequence (T1) (Figure 15, I–I'). During the earlier
See Figure 1 for location.
phase of base level lowering, erosion was followed by
deposition of continental sediments in the southern part
of the Lake Maracaibo basin. Deep marine conditions in Three sedimentary domains are recognized for
the northern part of the basin resulted in sedimentation sequence T1 (Figure 17). The first domain in the south-
of lowstand turbidites of the Trujillo Formation. In the western and southern parts of the Lake Maracaibo basin
second phase, transgression related to shelf flexure in is characterized by continental sedimentation of the
front of the applied nappe load reached the central part Mirador Formation. An inner shelf to shore zone domain
of the Lake Maracaibo basin. Toward the south, conti- occurs in the central Lake Maracaibo basin and is
nental deposition persisted, such as in the Mirador reflected in the sandstone-mudstone Misoa Formation.
Formation (Garner, 1926) (Figure 15, I–I'). Another trans- The third sedimentary domain in the north consists of
gression in the early Eocene deposited the stacked Misoa deep marine shales of the Trujillo Formation. The entire
“C” sandstones (sequences T1-1 to T1-5) (Figure 16). succession is up to 4000 m thick (Figure 17). Sedimenta-
Depositional sequence T1 culminated in a highstand tion rates vary from 190 m/m.y. in the shelf areas to 500
systems tract and deposition of the deltaic lower Misoa m/m.y. in the foredeep of the lateral ramp of the Lara
“B”(Sequence T1-6; Figure 16). nappes.
694 Parnaud et al.

Figure 17—Paleogeography of upper Paleocene–lower


Eocene depositional sequence T1. Legend: 1, Lara nappes,
actual position; 2, positive areas; 3, continental to deltaic
clastics; 4, inner to outer shelf sandstones and shales;
5, bathyal with turbidites, shales, and scarce sandstones;
6, thickness contours in feet.

This upper Paleocene–lower Eocene succession offers


significant petroleum exploration potential, with
numerous sandstone reservoirs in the transgressive and
regressive depositional systems. The deltaic intervals
require detailed analysis to characterize reservoir distrib-
ution.

Middle Eocene Sequences T2 and T3


During the middle Eocene, two major events changed
the configuration of the basin. First, southward encroach-
ment of the Lara nappes resulted in flexural subsidence
Figure 18—Eocene genetic stratigraphy. See Figure 1 for
of the Barinas-Apure basin shelf and in marine inunda- location.
tion. Basal sandstones of the Gobernador Formation
were followed by sedimentation of deep water shales of
the Pagüey Formation (Pierce, 1960) (Figure 18). Second, the upper Misoa “B” Formation (Figure 16), shelf flexure
tectonic loading by the Lara nappes produced a hinge resulted in turbiditic lowstand sedimentation and accu-
line along the Lake Maracaibo shelf in the northeastern mulation of the bathyal Paují Formation. Growth of the
sector. There, deposition of the shallow shelf upper Lara nappes was reflected in a forced northeastward
Misoa “B” sediments (Figure 16) was followed by deeper progradation (upper part of the Paují Formation).
water conditions and shale accumulations of the Paují After the highstand progradation of deltaic sediments
Formation (Tobler, 1922). This earlier sequence (T2) of the Cobre Formation in the southern part of the Barinas-
terminated in a progradational highstand system that Apure basin, a new transgressive cycle developed. These
was closely linked to the nappes and fed from the “Guanarito Sandstones” (Figures 15, J–J’, and 18) are ill
northeast instead of from the southwest (Figure 19). defined, but may benefit from study of the related Pagüey
Thus, following sedimentation of the basal sandstones of Formation which crops out in the Mérida Andes.
Stratigraphic Synthesis of Western Venezuela 695

Figure 19—Paleogeography of middle Eocene depositional Figure 20—Paleogeography of upper Eocene–Oligocene


sequences T2 and T3. Legend: 1, Lara nappes, actual depositional sequences T4 and T5. Legend: 1, Lara
position; 2, positive areas; 3, inner to middle shelf sand- nappes, actual position; 2, positive areas; 3, lacustrine to
stones and shales; 4, outer shelf to bathyal shales; 5, brackish sandstones, shales, and coal; 4, deltaic with
thickness contours in feet. marine influence, sandstones, and shales.

The thickness of these middle Eocene sequences T2 from Colombia (Carbonera Formation, Notestein, 1944;
and T3 ranges up to 1500 m. Estimated sedimentation and La Sierra Formation, Hedberg et al., 1937). Marine
rates are 20 m/m.y. in the shelf area and 150 m/m.y. in sediments were deposited in the eastern part of the basin
the foredeep. The hydrocarbon potential of these where it was open to the sea (Arauca Member of the
sequences is due to the thick sandstones bodies of the Guafita Formation; Ortega et al., 1987). The base of this
Misoa “C” (TST), Misoa “B” (TST and HST), and Gober- lower sequence corresponds seismically to an unconfor-
nador (TST) formations. mity (Figure 13A) that represents erosion of the Eocene
from west to east and erosion of the Paleocene section
south of Lake Maracaibo basin. There is seismic evidence
Upper Eocene–Lower Miocene that the Carbonera Formation pinches out. In the
Supersequence E: Collisional Basins Barinas-Apure basin, the top of this sequence is difficult
to map, although there are local onlap and truncation
Toward the end of the Eocene, the entire area changed boundaries. The sequence thins toward the north where
(Figure 2). Positive relief in the east and northeast it disappears (Kiser, 1989). To the south, near La Victoria
separated the continental Lake Maracaibo basin from the field, the geometry of this sequence is that of a transgres-
marine basin located in Falcón (Figure 20). Uplifts west sive system characterized by onlaps and a highstand
and south of the Serrania de Perijá and the eastern offlapping or progradational system.
Colombian Cordillera fed a fluviodeltaic depositional The second sequence (T5) was deposited in late
system. Marine circulation from the east continued to Oligocene–early Miocene time during widespread
affect the Barinas-Apure basin. This marine influence marine inundation (León Formation in Lake Maracaibo
extended to the Lake Maracaibo basin at the end of late basin, Notestein, 1944; and Guardulio Member of
Oligocene–early Miocene time. Guafita Formation in Barinas-Apure basin, Ortega et al.,
Two depositional sequences are recognized (Figures 3, 1987). The base of this sequence is unconformable and
20, 21). The first (T4) was deposited during the late marked by truncations and onlaps. The sequence thins
Eocene and early Oligocene in two different sedimentary toward the east of the Lake Maracaibo basin and wedges
domains. A deltaic domain in the western part was fed out along the Icotea structural trend (Figure 5, C–C’).
696 Parnaud et al.

Figure 22—Paleogeography of middle Miocene–Pleis-


tocene depositional sequence T7 related to Andean oro-
genesis. Legend: 1, Lara nappes, actual position; 2,
positive areas; 3, molassic depocenter; 4, lacustrine to
brackish sandstones and shales.

Rapid uplift was accompanied by molasse sedimenta-


tion along the margin of the Mérida Andes range.
Figure 21—Genetic stratigraphy showing comparison of
Cretaceous–Tertiary depositional sequences K5, T4, T5,
Marine sedimentation persisted in the Lake Maracaibo
and T6. See Figure 1 for location. basin, but it gradually changed to a freshwater paleo-
geography as the marine environment shrank toward
the north. In the Lake Maracaibo basin, a new transgres-
The thickness of these two sequences varies up to 1100 sive phase began during the middle Miocene and
m and has an estimated sedimentation rate of 50 m/m.y. resulted in deposition of the onlapping La Rosa
Sandstone horizons in this upper Eocene–lower Miocene Formation (Liddle, 1928). This was followed by regres-
succession have reservoir potential, including the Arauca sive progradation and contraction of marine influence
Member of the Guafita Formation and the Carbonera (Lagunillas Formation; Hedberg et al., 1937). Molasse
Formation. Local source rock intervals have been sediments of the Betijoque Formation (Garner, 1926)
described (Alberdi et al., 1994). were deposited along the Andean range at the same time
(Figure 22). Freshwater paleoenvironments dominated
Middle Miocene–Pleistocene the center of the Lake Maracaibo basin, as reflected in the
Supersequence F: Collisional Basins deposits of La Puerta (Garner, 1926 ) and Los Ranchos
(Liddle, 1928). Contemporaneous deposition of the
During the middle Miocene, large-scale compres- molasse Parángula and Río Yuca formations occurred in
sional tectonism initiated the Macizo de Santander, the Barinas-Apure basin (Mackenzie, 1937).
Serrania de Perijá, and Mérida Andes ranges. The These sequences locally exceed 5500 m in thickness in
Mérida Andes orogenesis culminated in the Plio-Pleis- the flexural foredeeps. Sedimentation rates greater than
tocene. This mountain building event correlates with two 250 m/m.y. have been estimated, but they decreased to
depositional sequences, T6 and T7 (Figure 3). Deforma- 150 m/m.y. along the foreland ramp. Several sandstone
tion also resulted in partition or isolation of the Lake intervals have reservoir potential, especially the Lagu-
Maracaibo and Barinas-Apure basins (Figure 22). Several nillas Formation. There are also subordinate reservoir
angular unconformities in the foothills of the northern intervals, such as the sandstones of the Santa Bárbara
and southern Andes record this tectonic history. Member in the La Rosa Formation.
Stratigraphic Synthesis of Western Venezuela 697

CONCLUSIONS collision. This forebulge migrated eastward ahead of the


advancing overthrust load in the Barinas-Apure basin
This study of the Lake Maracaibo and Barinas-Apure during the Paleocene.
basins has integrated several disciplines, including Integration of outcrop, well, and seismic data has
genetic stratigraphy from wells and outcrop sections and allowed us to resolve some problems. One such problem
seismic stratigraphy. Data acquired over many years by was the distribution of the Guafita Formation and its
different companies have been synthesized; 600 wells constituent Arauca and Guardulio members in the
were correleted and 4000 km of seismic lines interpreted. Barinas-Apure basin. Another example is the relation-
On this basis, stratigraphic and paleogeographic models ship between sedimentation of the different facies of the
have been developed that relate interaction among the Paují Formation and their geodynamic setting. The basal
Pacific, Caribbean, and South American plates and the part of the Paují Formation is composed of an outer shelf
way they periodically reorganized. to bathyal facies tract that reflects flexure of Lake
A new stratigraphic chart has been constructed in a Maracaibo basement due to the tectonic load of the
sequence stratigraphy framework. This analysis empha- nappes. The upper part of this formation is composed of
sizes the dynamics of basin subsidence and the tectonic more detrital facies related to a progradational highstand
control of sedimentation. Examples include the turbidite system fed from the nappes in the northeast.
facies of the Trujillo and Paují formations, which we Nevertheless, the stratigraphic columns of the
attribute to flexural subsidence of the foreland basin and Barinas-Apure and Lake Maracaibo basins require more
emplacement of the Lara nappes. Although tectonism work to standardize and interpret the stratigraphy. There
appears to have been the reason for the hierarchy of are several important stratigraphic issues that should be
unconformity-bounded depositional sequences, eustatic addressed. (1) The systems tracks and bounding uncon-
processes may have been important at times. However, it formities of the Peñas Altas Formation are yet to be
is not always possible to separate these components. resolved. (2) The precise age and distribution of the
The Mesozoic–Cenozoic geologic history of the study stratigraphic hiatus at the top of the Albian–lower Ceno-
area can be divided into several supersequences manian should be studied. (3) The evolution, diversity,
reflecting the history of extension and several episodes of and abundance of the late Cenomanian–early Campan-
collision. Six principal stages of basin formation corre- ian fauna requires attention. (4) Detailed work should
sponding to six supersequences were recognized and focus on the ages of the Eocene sequences. (5) The rela-
described. Sequence A accumulated in an extensional tionship of sedimentation of the Paují Formation to
setting. With the cessation of fault-controlled subsidence, emplacement of the Lara nappes along its lateral ramp is
passive margin supersequence B was deposited. not well understood, including definition of the lower
However, the presence of six depositional sequences transgressive component of the Lake Maracaibo shelf
(K0–K5) shows that postrift subsidence was episodic. and the upper progradational stratigraphy. (6) The strati-
This postrift margin includes the principal source rocks graphic relationships among the Oligocene sequences of
(e.g., La Luna Formation). The transition to a collisional the Lake Maracaibo and Barinas-Apure basins have not
margin and a flexural style of foreland basin subsidence been studied in detail.
occurred in supersequence C (three depositional
sequences, K6–K8). Supersequences D, E, and F reflect
multiple stages of compressional subsidence and
collision of the Caribbean and Pacific volcanic arcs. The Acknowledgments The authors wish to thank the manage-
Eocene Misoa and Gobernador formations (depositional ment of Intevep, S.A., for permission to publish this paper, as
sequences T1–T3) contain significant reservoir intervals. well as the industry sponsors of this project, Corpoven, S.A.,
Supersequence F also contains potential reservoir rocks Lagoven, S.A., and Maraven, S.A., the operational affiliates of
(e.g., La Rosa and Lagunillas formations). In this super- PDVSA, for financial support and for providing the subsurface
sequence, the Betijoque Molasse reflects rapid subsidence data. We also thank geologists François Roure and Bernard
along the orogenic front. Colletta (Institut Francais du Pétrole, France) for their
Estimated sedimentation rates reflect the dynamics of constructive criticisms of the manuscript. This paper has bene-
the tectonic and depositional environments. These sedi- fitted from the comments of Anthony Tankard, Anthony
mentation rates represent averages for the entire Edwards, and Ross McLean. We are grateful for the assistance
platform. The changes observed directly relate to of drafting personnel at Intevep, S.A.
changes in tectonic subsidence rather than environ-
mental conditions. The lowest rates correspond to
passive margin susidence (10–60 m/m.y.) and the
highest to the compressional phases, especially the time REFERENCES CITED
of encroachment of the Lara nappes (500 m/m.y.).
New biostratigraphy resolves several important strati- Alberdi, M., R. Tocco, and F. Parnaud, 1994, Análisis
graphic hiatuses (e.g., between the Lisure–Maraca and geológico integrado de las cuencas de Barinas y
the La Luna) in an area that includes the central and Maracaibo. Síntesis geoquímica de rocas: V Simposio Boli-
northern Perijá and a large part of the Lake Maracaibo variano, Exploración Petrolera de las Cuencas Subandinas:
basin. This particular hiatus is attributed to development Sociedad Venezolana de Geólogos, Caracas, Memoria,
of a forebulge associated with Pacific volcanic arc p. 411–412.
698 Parnaud et al.

Canache, M., A. Pilloud, I. Trustkowski, J. Crux, and S. Ortega, J. F., A. Van Erve, and Z. de Monroy, 1987, Formacion
Gamarra, 1994, Revisión estratigráfica de la sección Guafita: Nueva unidad litoestratigrafica del Terciario en el
cretácica del río Maraca, Serrania de Perijá, Venezuela: V subsuelo de la cuenca de Barinas-Apure, Venezuela suroc-
Simposio Bolivariano, Exploración Petrolera de las cidental: Boletin Sociedad Venezolana de Geologos, n. 10,
Cuencas Subandinas, Sociedad Venezolana de Geólogos, 23 p.
Caracas, Memoria, p. 240–241. Parnaud, F., I. Truskowski, Y. Gou, M. A. Capello, B. de Toni,
Christ, P., 1927, La coupe geologique le long du chemin de J.-C. Pascual, A. Sanchez, A. Pilloud, M. Canache, and S.
Mucuchachi a Santa Barbara dans les Andes Venezueli- Gamarra, 1994, Modelo biolitoestratigráfico del Occidente
ennes: Eclogae Geologicae Helvetiae, v. 20, p. 397–414. de Venezuela: V Simposio Bolivariano, Exploración
Garner, A. H., 1926, Suggested nomenclature and correlation Petrolera de las Cuencas Subandinas, Sociedad Vene-
of geological formations in Venezuela: American Institute zolana de Geologos, Caracas, Memoria, p.161–163.
of Mining and Metallurgical Engineers Transactions, Pierce, G. R., 1960, Geología de la cuenca de Barinas:
p. 677–684. Congreso Geologico Venezolano III, Caracas, 1959, t. 1,
González de Juana, C., J. M. Iturralde de Arozena, and X. p. 214–276.
Picard Cadillat, 1980, Geologia de Venezuela y de sus Pindell, J. L., and J. P. Erikson, 1993, The Mesozoic passive
cuencas petrolíferas: Caracas. Ediciones Foninves, v. 2, margin of northern South America, in A.Vogel, ed., Creta-
1051 p. ceous tectonics in the Andes: International Monograph
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of Series, Earth Evolution Sciences, Wiesbaden, FRG, Vieweg
fluctuating sea levels since the Triassic: Science, v. 235, Publishing, p. 1–30.
p. 1156–1167. Renz, O., 1959, Estratigrafía del Cretáceo en Venezuela Occi-
Hea, J. P., and A. B. Whitman, 1960, Estratigrafia y petrologia dental: Boletin Geologico, Caracas, v. 5, p. 3–48.
de los sedimentos precretacicos de la parte norte-central de Rod, E., and W. Maync, 1954, Revision of Lower Cretaceous
la Serrania de Perija, Estado Zulia, Venezuela: Conferencía stratigraphy of Venezuela: AAPG Bulletin, v. 38,
Geológica Venezolana, Caracas, Publicación Especial 3, p. 193–282.
p. 351–376. Sievers, W., 1888, Die Kordilliere von Merida nebst
Hedberg, H. D., 1931, Cretaceous limestones as petroleum Bemerkungen uber das Karibische Gebirge: Geogogische
source rocks in northwestern Venezuela: AAPG Bulletin, v. Abhandlungen Hessen (Penck), v. 3, 238 p.
15, p. 229–244. Stainforth, R. M., 1962, Definitions of some stratigraphic units
Hedberg, H. D., and L. C. Sass, 1937, Sinopsis de las forma- in western Venezuela: Las Pilas, Cocuiza, Vergel, El Jebe,
ciones geológicas de la parte occidental de la Cuenca de Tres Esquinas, and Nazaret: Asociación Venezolana de
Maracaibo, Venezuela: Caracas, Boletin Geológia y Minas, Geólogos Mineros y Petroleros, Boletin, v. 5, p. 279–282.
t. 1, p. 77–120. Stephan, J. F., 1977, Andes et Chaine Caraibe sur la transver-
Hodson, F., 1926, Venezuelan and Caribbean turritellas, with sale de Barquisimeto,Venezuela: Evolution Géody-
a list of Venezuelan type stratigraphic localities: American namique,Géodynamique des Caraibes, Symposium, Paris,
Paleontology Bulletin, v. 11, p. 173–220. Edition Technip, p. 505–529.
Kehrer, L., 1938, Algunas observaciones sobre la estratigrafía Sutton, F.A., 1946, Geology of the Maracaibo basin,
en el Estado Táchira: Caracas, Boletin Geológia y Minas, t. Venezuela: AAPG Bulletin, v. 30, p. 1621–1741.
2, p. 44–56. Tobler, A., 1922, Die Jacksontufe (Priabonien) in Venezuela
Kiser, G. D., 1989, Relaciones estratigraficas de la cuenca und Trinidad: Eclogae Geologicae Helvetiae., v. 17,
Apure-Llanos con areas adyacentes, Venezuela suroeste y p. 342-346.
Colombia Oriental: Caracas, Sociedad Venezolana de Zambrano, E., E. Vasquez, B. Duval, M. Latreille, and B.
Geólogos, v. 1, p. 71. Coffinieres, 1970, Synthése paléogéographique et
Künding, E., 1938, Las rocas pre-cretaceas de los Andes pétroliere du Venezuela occidental: Revue de l’Institut
centrales de Venezuela con algunas observaciones sobre su Français du Pétrole, Paris, Edition Technip, v. 25, p.
tectónica: Boletin Geológia y Minas, Caracas, t. 2, p. 21–43. 1449–1491.
Liddle, R. A., 1928, The geology of Venezuela and Trinidad:
Fort Worth, Texas, J. P. MacGowan, 552 p.
Liddle, R. A., G. D. Harris, and J. W. Wells, 1943, The Rio Authors’ Mailing Addresses
Cachiri section in the Sierra Perijá, Venezuela: American
Paleontology Bulletin, v. 22, p. 273–365. François Parnaud
Mackenzie, A. N., 1937, Sección geológica de la región de Yves Gou
Barinas: distritos Barinas, Bolivar y Obispos del Estado Jean-Claude Pascual
Barinas, Venezuela: Caracas, Boletin Geológia y Minas, t.1, Beicip-Franlap
p. 269–283. Petroleum Consultants
Martín Bellizzia, C., 1968, Edades isotopicas de rocas vene- B.P. 213
zolanas: Boletin Geológia, Caracas, v. 9, p. 356–380. 92500 Rueil-Malmaison
Menéndez, V. de V., A., 1968, Revisión de la estratigrafía de la France
Provincia de Pastora según el estudio de la región de
Guasipati, Guyana venezolana: Caracas, Boletin Geológia,
v. 9, p. 309–338.
Maria Angela Capello
Notestein, F. B., C. W. Hubman, and J. W. Bowler, 1944, Irene Truskowski
Geology of the Barco concession, Republic of Colombia, Herminio Passalacqua
South America: GSA Bulletin, v. 55, p. 1165–1216. Intevep, S.A.
Oppenheim, V., 1937, Contribución a la geología de los Andes Apartado Postal 76343
Venezolanos: Caracas, Boletin Geológia y Minas, v. I, Caracas 1070A
p. 25–45. Venezuela
Two-Dimensional Computer Modeling of Oil
Generation and Migration in a Transect of the
Eastern Venezuela Basin

Oswaldo Gallango François Parnaud


Intevep, S.A. Beicip-Franlab
Caracas, Venezuela Rueil-Malmaison, France

Abstract

T he purpose of this two-dimensional computer simulation of basin evolution, based on geologic,


geophysical, geochemical, geothermal, and hydrodynamic data, was to determine the hydrocarbon
generation and migration history of the basin. The modeling covered two geologic sections (platform and
prethrusting) located along the Chacopata-Uverito transect in the Eastern Venezuela basin. In the platform
section, a hypothetical source rock equivalent to the Guayuta Group was used to simulate the migration of
hydrocarbons. The thermal history reconstruction of the hypothetical source rock confirms that it would not
have reached the oil window before the middle Miocene and that the maturity in this sector is due to the sedi-
mentation of the Freites, La Pica, and Mesa–Las Piedras formations. The expulsion of the hydrocarbons took
place mainly into the Oligocene–Miocene reservoir and has not yet reached zones located beyond the
Oritupano field. This implies that the oil in the southern part of the basin was generated by a source rock
located to the north, in the actual deformation zone. For the past 17 m.y., water has migrated from north to
south in this section. In the prethrusting section, the hydrocarbon expulsion started during the early Tertiary
and migrated mainly into Lower Cretaceous reservoirs (El Cantil and Barranquín formations). At the end of
the passive margin stage, hydrocarbons migrated across the Merecure reservoir zone and into the Onado area
before thrusting began.

Resumen

E l propósito del presente estudio fue simular en un modelo computarizado en dos dimensiones, la
evolución de una cuenca sobre la base de datos geológicos, geofísicos, geoquímicos, geotérmicos e
hidrodinámicos, con la finalidad de establecer la historia de generación y migraciòn de los hidrocarburos. El
modelaje abarca dos secciones geológicas (plataforma y pre-corrimiento) localizadas a lo largo del transecto
Chacopata-Uverito de la Cuenca Oriental de Venezuela. En la sección de plataforma una roca madre hipotética
equivalente al Grupo Guayuta fue considerada con la finalidad de simular la migración de hidrocarburos. La
reconstrucciòn de la historia termal de la roca madre hipotética, confirma que esta no alcanzó la ventana de
petróleo antes del Mioceno Medio y que la madurez en este sector es debida a la sedimentación de la forma-
ciones Freites, La Pica y Mesa-Las Piedras. La expulsión de los hidrocarburos tuvo lugar principalmente hacia
el yacimiento Oligo-Mioceno y no alcanzó zonas localizadas más alla del campo Oritupano. Esto implica que el
petróleo en la parte sur de la cuenca fue generado por una roca madre localizada en la parte norte, en la actual
zona de deformación. En los últimos 17 m.a. ,el agua ha estado migrando de norte a sur en esta sección. En la
sección pre-corrimiento la expulsión de hidrocarburos comenzo durante el Terciario temprano y migro princi-
palmente hacia los yacimientos del Cretácico temprano (formaciones El Cantil y Barranquín). A finales del
margen pasivo, los hidrocarburos migraron a través del yacimiento Merecure hasta la zona de Onado, antes de
que comenzaran los corrimientos.

Gallango, O., and F. Parnaud, 1995, Two-dimensional computer modeling of oil genera- 727
tion and migration in a transect of the Eastern Venezuela basin, in A. J. Tankard, R.
Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62,
p. 727–740.
728 Gallango and Parnaud

aim of the program is to help the explorationist with the


vertical and horizontal integration of the physical and
chemical parameters that influence the evolution of a
sedimentary basin. The principal parameters involved in
this type of modeling are as follows:

• Stratigraphy, including lithologies of basin deposits


• Temperature and heat flow
• Pressures and overpressures associated with the
basin
• Geochemical characteristics of the source rock and
fluids

The program consists of several modules. Backstrip-


ping is used to study the geometric evolution of the
basin. This run is normally done using simple com-
paction via a porosity-depth law, an imposed tempera-
ture gradient, and no geochemical data. A run is used to
calibrate the calculated data with the known data from
Figure 1—Location of the Chacopata-Uverito transect
situated in the western part of the Eastern Venezuelan the basin, particularly the distribution of reservoir
basin. The 250-km-long section (shaded) runs across the porosity, pressure and overpressure associated with
Maturin subbasin from north to south. those reservoirs, and water flow. These parameters are
lithology dependent. In this study, effective stress–
porosity relationships were reconstructed from wireline
INTRODUCTION logs and pressure data (RFT, mud density). This module
was used to simulate sedimentary rock compaction with
This integrated study of the physical and geochemical a single fluid phase (always water) using an imposed
phenomena that occur during the evolution of a sedi- temperature gradient and without geochemical data. A
mentary basin (compaction, fluid movement, heat thermal module was used to calibrate the distribution of
transfer, generation, and migration of hydrocarbons) the known temperatures in the basin. The TEMISPACK
through the use of a numerical model opens a new era in program allows an imposed geothermal gradient or heat
hydrocarbon exploration. These models reconstruct the flow. This calculation was done with one fluid phase
geometrical evolution of a basin in two dimensions, as (water) and without geochemical data.
well as the fluid movement patterns, paleopressure, A maturity run calibrates the maturity observed in the
paleotemperature, hydrocarbon potential, and accumula- basin through laboratory geochemical analysis (Tmax,
tion throughout geologic time up to the present. These % Ro) with maturity calculated by TEMISPACK. This
models are widely used as an exploratoration guide, run is done with only one phase (water) using the
especially in deep and frontier basins that require maturity of the source rock without migration of fluids
analysis of all available geologic information in an and with an imposed temperature gradient or heat flow,
attempt to reduce exploration risk. This is why these according to thermal module results. Several module
computerized methods of predicting possible hydro- passes are made to test sensitivity. The calculations are
carbon accumulations are becoming increasingly done with two fluid phases (water and hydrocarbons)
important. and a temperature gradient or heat flow, according to
This paper presents the results of TEMISPACK inte- thermal model results and the following three options:
grated numerical modeling of the petroleum generation (1) without secondary cracking, to obtain migration of all
and migration history along two north-south regional expelled hydrocarbons as a unique fluid phase and
cross sections located in the western part of the Eastern without any significant phase change during migration;
Venezuela basin (Figure 1). The purpose of the study is (2) with secondary cracking, to obtain petroleum
to establish the time when expulsion of hydrocarbons migration only (a convenient way to evaluate the timing
began and to which reservoir units they migrated; and to of oil migration and the conditions for preserving oil);
establish the transformation rate of the kerogen into and (3) with primary and secondary cracking, to obtain
hydrocarbons and the remaining oil potential just before gas migration (methane to C5). The oil is cracked in the
the thrusting began. This modeling is part of a regional source rock prior to any expulsion.
synthesis carried out by Intevep.

The TEMISPACK Program MODELING OF THE PLATFORM SECTION


TEMISPACK is a finite volume model that recon- Parameters
structs the history of petroleum generation and migra-
tion along a two-dimensional evolutionary mesh repre- The platform section (Figure 2) was simulated to
senting a regional cross section (Burrus et al., 1992). The calibrate the computerized data against the real well
Figure 2—Generalized north-south structural cross section (Orinoco River to Cumaná) showing the two sections
modeled in this study. (From Parnaud et al., 1995.)
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin
729
730 Gallango and Parnaud

Figure 3—Generalized
stratigraphy of the study
area showing basin types,
source rocks, and reser-
voirs. (From Parnaud et al.,
1995.)

data, including the following: in the south. This excess pressure diminishes until it
disappears into the Cerro Negro area (Figure 6), where
• Porosity in the principal reservoirs: the the Oligocene–Miocene reservoir is directly in contact
Oligocene–Miocene reservoir (Merecure and with the main upper water-bearing zone (Mesa–Las
Oficina formations) and the upper sandstones of Piedras).
the Mesa and Las Piedras formations (Figure 3)
• Known pressures in both main reservoirs:
Mesa–Las Piedras sandstones and Oligocene– Modeling Results
Miocene reservoirs and associated overpressures
• Aquifer characteristics: Mesa–Las Piedras sand- Several runs were carried out to calibrate the calcu-
stones are the present aquifers in the basin and are lated data against the observed pressure, overpressure,
recharged from the north and discharged into the porosity, temperature, organic maturity, and other
Orinoco River to the south thermal indices. Calculation of pressure was done on the
basis of several hypotheses related to north and south
The existence of a potential source rock is unknown in end-members, with and without active water flow.
the platform section, unlike in the north where the The boundary conditions that best explain the present
Querecual and San Antonio formations of the Guayuta conditions are an open aquifer in the southern area
Group are believed to be prolific. However, to simulate (Orinoco River) and a closed aquifer in the northern area
migration, a hypothetical source rock was assumed at the just south of the deformation front. This allows the
extreme north of the section (Figures 4, 5). Points of observed overpressures in the Oligocene–Miocene
reference and porosities for the platform Mesa–Las reservoir and the almost hydrostatic pressure in the
Piedras, Oficina, and Merecure formations are show in Mesa–Las Piedras sandstones to be simulated. As seen in
Table 1. For reference, pressure data were taken from Figure 7, there is excellent correlation between the calcu-
existing measurements between the main upper aquifer lated and observed overpressures in the sandstones of
(Mesa–Las Piedras sandstones) and the Oligocene– the Oligocene–Miocene reservoir in the northern area of
Miocene reservoir levels of the Merecure and Oficina the section, including the near hydrostatic pressure in the
formations ( Figure 6). Mesa–Las Piedras sandstones, as well as the Miocene
In the phreatic, water-bearing Mesa and Las Piedras sandstones in the southern section. In Figure 8, the actual
formations, a slightly higher pressure than the hydro- distribution of overpressures can be observed to
static pressure is noticed. The differential pressure diminish toward the south where there is contact and
reflects the altitude above sea level of the northern equilibrium between the Miocene sandstones and the
recharge zone; pressure decreases progressively until water-bearing Mesa and Las Piedras formations. The
discharge into the Orinoco River. In the Oligocene– porosity calculation done with the TEMISPACK pro-
Miocene reservoir located in the northern area near the gram was verified at several control points containing
deformation front, an important difference in pressure information about the two main reservoirs (Mesa–Las
occurs that is related to tectonic loading by thrust sheets Piedras and Oligocene–Miocene) (Table 1). In Figure 9, a
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin 731

Figure 4—Stratigraphy of the platform section.

Figure 5—Present day regional cross section showing the distribution of the seven lithofacies considered in the models.
732 Gallango and Parnaud

Table 1—Formation Porosity Values at Reference Points

Porosity (%)
Mesa–
Sector Las Piedras Oficina Merecure
Onada (km 120) 35 11 9
Oritupano (km 80) 35 15 17
El Salto (km 55) 35 25 29
Cerro Negro (km 30) 35 31 35
South Cerro Negro (km 17) 35 35 35

good correlation between calculated values and


measured values from the wells can be observed.
The water flow study shows vertical migration in low
permeability shales and horizontal migration in the
reservoir levels. Starting about 17 m.y. before thrusting,
water has migrated from north to south toward its
confluence with the Orinoco River (Figure 10). Water
flow is weak in the Oligocene–Miocene reservoir, with
an average speed of 1 km/m.y., but it reaches a much
higher speed at the upper water-bearing Mesa and Las
Piedras formations. Flow rates as high as 1500 km/m.y.
have been documented where the aquifer discharges into
the Orinoco River (Figure 10).
Thermal calibration involved several working
hypotheses in an attempt to calibrate these characteristics
with the maturity data available in wells. A good correla-
tion between simulated and measured well temperatures
was obtained assuming a gradient of 28°C/km or a heat
flow of 46 mW/m2. However, these two values give
extremely high maturity levels, and in order to establish
a good calibration with the maturity obtained from the
organic matter in the wells, a gradient of 22°C/km was
used. A lower heat flow of 42 mW/m2 was derived from
the first run of the simulations.
The historical reconstruction of maturity confirms that
Figure 6—Differentiated pressure zones along the transect
the hypothetical source rock would not have reached the according to original pressure values.
oil window before middle Miocene time (postthrusting)
and that maturity is attributed to Freites, La Pica, and
Mesa–Las Piedras sedimentation. Maturity levels at
present are predicted to be very high (Ro > 3.6%), which
implies that the hypothetical source rock should have
generated and expelled hydrocarbons completely
(Figure 11).
Simulation with a hypothetical source rock laterally
equivalent to the Guayuta Group has several implica-
tions. (1) Expulsion and migration of hydrocarbons
occurred after the middle Miocene, mainly during sedi-
mentation of the La Pica and Mesa–Las Piedras forma-
tions. Expulsion of hydrocarbons took place toward the
reservoir sandstones of Merecure, Oficina, and Lower
Cretaceous formations. (2) At present, migrating hydro-
carbons do not reach as far as the Oritupano area (Figure
12), which suggests that previous migration from the
northern part of the basin located in the deformation belt
must be invoked to explain the occurrence of petroleum
in the south. (3) Migration of hydrocarbons occurred
mainly within the Oligocene–Miocene sandstones Figure 7—Profile showing a comparison of observed and
(Merecure and Oficina formations); however, important computed overpressure (P) values.
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin 733

Figure 8—Computed overpressure distribution (in MPa) at the present time for the platform section. The white area repre-
sents the region of hydrostatic pressure.

saturation occurs in the lowest Cretaceous and older.


(4) An average gradient of 27.2°C/km observed in the
wells varies from 36°C/km in the south to 22°C/km in
the deepest part of the section to the north.

MODELING OF THE PRETHRUSTING


SECTION
Parameters
The prethrusting section was modeled to simulate
hydrocarbon generation and its migration toward the
reservoirs in the southern part of the basin prior to
thrusting in the Oligocene–early Miocene. The model
addresses the structural restoration as well as the strati- Figure 9—Relationship between porosity and depth in the
graphic reconstruction obtained from the geologic model Oritupano area.
of the transect. In a simulation of this kind, classic cali-
bration with known data in the basin such as porosity
and pressure is not possible because of lack of age stratigraphic model of the Chacopata-Uverito transect, the
constraints. Nevertheless, to overcome this limitation, a eroded section was estimated in order to reconstruct the
section in the nondeformed area of the basin was prethrusting section (Figures 3, 13). A reconstruction of the
simulated from the deformation front to the Orinoco northern part of Turimiquire was not undertaken because of
River, where a good calibration of parameters was the complete absence of surface and subsurface geologic
achieved. Such parameters were used for the pre- information due to erosion of the Cretaceous section up to
thrusting section, and especially for the porosity–depth the Barranquín Formation; this section undoubtedly
data from the Merecure reservoir. A normal passive contributed to hydrocarbon generation. To simulate this
margin type basin without overpressure was modeled. section with TEMISPACK, a lower Miocene top has been
Maturity data from several wells along the reconstructed used as a reference level and an arbitrary age of 10 Ma
section permitted a heat flow calibration. From the litho- assigned to it.
734 Gallango and Parnaud

Figure 10—Water flow (in m/m.y.) in the platform section at the present time.

Figure 11—Vitrinite reflectance (% Ro) distribution in the platform section at the present time.
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin 735

Figure 12—Hydrocarbon saturation (%) in the platform section at the present time. Arrows indicate the principal direction of
hydrocarbon migration.

Figure 13—Reconstructed stratigraphy in the prethrusting section at the end of early Miocene time.
736 Gallango and Parnaud

Figure 14—Reconstructed facies distribution in the prethrusting section at the end of early Miocene time.

Stratigraphic Scale of the Section (a)


Figures 3 and 13 show the stratigraphic scale of the
section. On seismic lines, pre-Cretaceous levels were
assigned to the Jurassic and could reach a significant
thickness (Figure 14). It was randomly divided into four
units that were each assigned a theoretical facies
consisting of a sandstone-shale mix and taking into
account all relevant data from the area (e.g., Espino
graben). As reference, results of platform sections were
used, as well as laboratory and porosity-depth data from
Oligocene–Miocene formations. Pressure data are not
available for calibration for the end of the early Miocene,
but there is no evidence that overpressure was present at
that time. Nevertheless, below 3–4 km, the existence of
overpressure can be anticipated because of the thickness.
The only parameter available to calibrate this section
is the maturity level observed in wells located in the area.
The data assigned to four models or wells (F0–F3)
(Figure 2) were integrated, and the following results (b)
were observed.

Well F0
TEMISPACK and laboratory data do not agree
because much higher maturity values were obtained
from the laboratory data. The results shown in Figure
15a suggest that at the end of the passive margin stage
and just before the onset of thrusting, the sedimentary
column (Querecual–lower Carapita) and the associated
organic matter had not reached the maturity level shown
by these formations today. This suggests that the
maturity level observed at present must have been
acquired after the passive margin stage and that it was
associated with tectonic and sedimentary subsidence Figure 15—Reconstructed vitrinite reflectance (% Ro) distri-
(Miocene–Pleistocene deposits, La Pica, and Mesa–Las bution in the prethrusting section at the end of early
Piedras). Miocene time for (a) well F0 and (b) well F3.
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin 737

Petroleum density in the subsurface is 30% lower than


at the surface (England et al., 1987), and the low maturity
of slightly biodegraded oil from Oritupano area has a
surface density of 920 kg/cm3 (22° API) (Gallango et al.,
1992). Because of these two factors, a value of 644
kg/cm3 for the oil density in the subsurface was taken
for migration purposes.

Modeling Results
As for the platform section, several models were
performed for the prethrusting section with various
parameters. A porosity-depth law was assumed for the
Merecure Formation on the basis of what is already
known for the platform zone. A normal pressure distrib-
Figure 16—Reconstructed vitrinite reflectance (% Ro) distri- ution and a heat flow of 42 mW/m2 calibrated for the
bution in the prethrusting section at the end of early
platform section were used.
Miocene time for well F1.
In the south, from El Furrial to the southern limit of
the source rock (Figure 15a), maturity was attained after
Wells F1 and F2 thrusting and was associated with either tectonic
loading or sedimentation prior to the early Miocene.
Simulated and laboratory results indicate an excellent
This sedimentation includes the upper Carapita, Freites,
correlation for both vitrinite reflectance values and Tmax
La Pica, and Mesa–Las Piedras formations. The maturity
(Figure 16). The later implies that present maturity levels
level reached at the end of early Miocene time matches
from the studied formations were established before
the present maturity and was acquired before thrusting
thrusting and during passive margin subsidence and,
in the area corresponding to the Pirital sheet (Figure 16).
consequently, that the maturity of the Pirital sheet section
In the Serranía del Interior, the high maturity level was
was acquired in the early Miocene.
acquired after the early Miocene and may be related to
Well F3 tectonic loading due to thrusting (Figure 15b). Figure 17
There is no correlation between laboratory and shows the maturity level reached at the end of the early
simulated maturity (Figure 15b) in this well. The present Miocene throughout the simulated section.
maturity level varies between 1.5 and 2% Ro, while To study hydrocarbon migration, a simulation with-
maturity values reached at the end of the early Miocene out secondary cracking was undertaken using the low
(calculated by TEMISPACK) were no greater than 1.5% gas–oil ratio observed over the entire area (356 m3/m3).
Ro and varied between 1.0 and 1.5% Ro. On this basis, it Saturation resulting from maturity, expulsion, and
is concluded that the sedimentary column and the associ- migration of hydrocarbons leads to several conclusions.
ated organic matter attained this high maturity level after Hydrocarbon expulsion began during the early Tertiary
the passive margin stage due to the sedimentation of a (Figure 18). At first, expulsion took place toward the
postorogenic formation (the Naricual Formation) or by Lower Cretaceous reservoir through the El Cantil
the tectonic emplacement and loading of thrust sheets. It Formation (Figure 18). Migration and expulsion toward
is possible that a higher geothermal gradient occurred the Merecure reservoir took place at the beginning of the
during the development of the passive margin. early Miocene (Figure 19). At the end of the passive
The Guayuta Group (Querecual and San Antonio margin stage, the principal migration invaded the entire
formations) was identified as a source rock with marine Merecure reservoir and may have reached the Onado
organic matter (type II). This source rock extends from area. At lower levels (Barranquín Formation), oil satura-
the southern edge of the Pirital thrust (177 km) to the tion is high. The supposed Jurassic deposit was not
northern part of the section (Figures 13, 14). It consists of saturated by hydrocarbons. The transformation ratio of
shales and pelagic limestones. The present maturity the source rock varied from 10 to 55% at the end of the
levels for the Guayuta Group are variable, with vitrinite early Miocene (Figure 20).
reflectance values of 0.6% Ro being observed in the The prethrusting section model produced several
southern zone and increasing to 2.0 % Ro in the Serranía interesting results. The expulsion of hydrocarbons took
del Interior (R o > 1.3%). With this extremely high place preferentially toward the reservoir levels of the El
maturity level, the present average TOC ranges from 2.0 Cantil and Barranquìn formations. In a second stage,
to 6.0 wt. %, which indicates that the original TOC was migration took place toward the Merecure reservoir
higher than 6.0 wt. %. through which hydrocarbons migrated before thrusting
TEMISPACK modeling and high TOC values suggest started in the Onado area. The transformation ratio of the
an oil potential higher than 20 kg HC/t of rock. Consid- source rock at the end of the early Miocene was in the
ering that the northern area was not simulated because range of 10–55%, which implies that an important
of limited information, very high values of S2 (80 kg/t) remnant potential generator existed for the period
have been estimated to overcome previous limitations. during and after thrusting.
738 Gallango and Parnaud

Figure 17—Maturity (% Ro) of organic matter in the prethrusting section at the end of early Miocene time.

Figure 18—Hydrocarbon saturation (%) in the prethrusting section at the end of Oligocene time.
Two-Dimensional Computer Modeling of Oil Generation and Migration, Eastern Venezuela Basin 739

Figure 19—Hydrocarbon saturation (%) in the prethrusting section at the end of early Miocene time.

Figure 20—Computed transformation ratio in the prethrusting section at the end of early Miocene time.
740 Gallango and Parnaud

CONCLUSIONS England, W., A. Mackenzie, D. Mann,. and T. Quigley, 1987,


The movement and entrapment of petroleum fluids in the
Computer simulations of oil generation and migration subsurface: Journal of the Geological Society of London, v.
in the Chacopata-Uverito transect have been carried out 144, p. 327–347.
using available porosity, pressure, and maturation data Gallango, O., M. Escandòn, M. Alberdi, F. Parnaud, and J.-C.
as constraints. The oil expulsion from the Cretaceous Pascual, 1992, Hydrodynamism, crude oil distribution, and
geochemistry of the stratigraphic column in a transect of
Querecual and San Antonio source rocks located in the the Eastern Venezuelan basin (abs.): GSA 29th Annual
deformation zone (Pirital sheet and Serranìa del Interior) Meeting, Cincinnati, Ohio, Abstracts with Programs,
began during the early Tertiary. Expulsion was focused p. A214.
mainly toward the Lower Cretaceous (El Cantil and Parnaud, F., Y. Gou, J.-C. Pascual, I. Truskowski, O. Gallango,
Barranquìn formations) and secondarily across the H. Passalacqua, and F. Roure, 1995, Petroleum geology of
Merecure reservoir, through which the oil migrated the central part of the Eastern Venezuelan basin, in A. J.
progressively southward before thrusting. The oil in the Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins
southern part of the basin corresponds to a first genera- of South America: AAPG Memoir 62, this volume.
tion that occurred in the Serranía del Interior rather than
in a hypothetical source rock located in the proximity of
the deformation front.

Acknowledgments This paper includes the results of work


financially supported by Lagoven, S.A., Corpoven, S.A., and
Maraven, S.A. The authors are grateful to PDVSA and
Authors’ Mailing Addresses
Intevep, S.A., for authorizing the presentation and publication
of this paper. Special thanks are due to Jim Allan and Anthony Oswaldo Gallango
Tankard for critical reading and suggesting improvements. Intevep, S.A.
Apartado de correos 76343
Caracas 1070-A
References Cited Venezuela

Burrus, J., E. Brose, G. de Janvry., Y. Grosjean, and J. Oudin, François Parnaud


1992, Basin modelling in the Mahakan Delta based on the Beicip-Franlab
integrated 2D model TEMISPACK: Twentieth Annual B.P. 213
Convention, Proceedings of the Indonesian Petroleum 92500 Rueil-Malmaison
Association, Jakarta, 20 p. France
Petroleum Geology of the Central Part of the
Eastern Venezuelan Basin

François Parnaud Irene Truskowski


Yves Gou Oswaldo Gallango
Jean-Claude Pascual Herminio Passalacqua
Beicip-Franlab Intevep, S.A.
Rueil-Malmaison, France Caracas, Venezuela

F. Roure
IFP
Rueil-Malmaison, France

Abstract

T wo main petroleum provinces are described for the central part of the Eastern Venezuelan basin. These
include the southern foreland platform near the Orinoco, with its heavy oil fields of the Faja Petrolifera,
and in the north, the giant El Furrial and related traps in the frontal thrusts of the Serrania. To identify the
petroleum systems and define attractive petroleum plays in this complex foreland fold and thrust belt system,
an integrated geologic model of the area has been built using stratigraphic, structural, reservoir, and geochem-
ical data. The major results of the study are as follows: (1) a reliable stratigraphic synthesis including updated
ages; (2) a coherent structural interpretation in the thrusted zones validated by a balanced north-south cross
section, with relevant new interpretations being the discovery of out-of-sequence thrusts (Pirital thrust) and the
involvement of upper crust in the deformation; (3) the description of reservoir and hydrodynamic models that
emphasize the importance of the major Las Piedras, Oficina, and Merecure reservoirs; and (4) characterization
of three different source rocks of marine and continental origin in the Guayuta Group (two marine sources)
and Carapita Formation (one continental source).
To better understand the petroleum generation in this area—the maturation of potential source rocks,
expulsion, and migration of the hydrocarbons—one- and two-dimensional numerical modeling has also been
attempted in the platform area for times before and after thrusting. This method has given an estimate of
420–1350 billion bbl of generated oil, of which 6–15% have been recognized within the study area.

Resumen

S e describen dos provincias petrolíferas principales en la parte central de la Cuenca Oriental de


Venezuela: En la plataforma de antepaís cercana al Orinoco hacia el Sur, los campos de petróleo pesado
de la Faja Petrolífera; y al Norte el campo gigante El Furrial y las trampas relacionadas en los cabalgamientos
frontales de la Serranía del Interior. Para identificar los sistemas petroleros y definir yacimientos petrolíferos
atractivos en este complejo sistema de cinturón de pliegues y cabalgamientos-cuenca de antepaís, un modelo
geológico integrado del área ha sido construido por primera vez usando datos estratigráficos, estructurales,
geoquímicos y de yacimiento. Los principales resultados del estudio se refieren a: (1) la elaboración de una
síntesis estratigráfica confiable que incluye dataciones actualizadas; (2) una interpretación estructural coherente
en las zonas de cabalgamiento, validada por el balanceamiento de una sección transversal Norte-Sur; nuevos
hechos relevantes son el descubrimiento de cabalgamientos fuera de secuencia (el Cabalgamiento de Pirital) y
la implicación de la corteza superior en el proceso de deformación; (3) la descripción de modelos de yacimiento
e hidrodinámico, que revelan la importancia de los yacimientos principales Las Piedras, Oficina y Merecure;
(4) una caracterización de tres diferentes rocas madre de origen marino y continental en el Grupo Guayuta (de
origen marino) y la Formación Carapita (de origen continental).
Para mejorar el entendimiento de la generación de petróleo en esta área, es decir, la maduración de las rocas
madre potenciales, la expulsión y migración de los hidrocarburos, se probaron modelajes numéricos 1-D y 2-D
en el área de plataforma antes y después de los cabalgamientos. De esta manera, se ha estimado una cantidad
entre 420–1350 miles de millones de barriles de petróleo generado, de los cuales, de 6–15 por ciento se han
encontrado dentro del area estudiada.

Parnaud, F., Y. Gou, J.-C. Pascual, I. Truskowski, O. Gallango, H. Passalacqua, and F. Roure, 1995, 741
Petroleum geology of the central part of the Eastern Venezuela basin, in A. J. Tankard, R. Suárez S.,
and H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 741–756.
742 Parnaud et al.

INTRODUCTION the belt (Carnevali, 1988a) (Figure 1).


Numerous studies, some of them addressing the
Bordered to the north by a fold and thrust belt—the stratigraphy (Hedberg, 1950; Sulek, 1961; Arnstein et al.,
Serrania del Interior—and by the El Pilar fault, which 1985; Chiock, 1985; Parnaud et al., 1992a), and others
mark a major plate boundary with the Caribbean, the based on seismic data (Rossi, 1985; Rossi et al., 1985;
Eastern Venezuelan basin extends as far south as the Carnevali, 1988b; Potié, 1989; Parnaud et al., 1992b;
Orinoco River (Hedberg, 1950). This basin is a flexure- Roure et al., 1994), have already described the major
related foreland basin that developed in Neogene time regional tectonic features and large oil fields. However, a
on the late Mesozoic passive margin of the South synthesis integrating all the available surface and subsur-
American craton. It comprises two of the major face data still remains to be done.
petroleum provinces of this continent (Gonzalez de Extensive seismic surveys have been acquired and
Juana et al., 1980): (1) the extensive fields of the Faja numerous exploration wells have been drilled in the
Petrolifera in the southern foreland platform near the central portion of the basin (Figure 1). This unique set of
Orinoco River (Fiorillo, 1982) (Figure 1), and (2) the data is used here, together with new biostratigraphic and
surface seeps and shallow prospects of the Quiriquire geochemical analyses, structural studies, and numerical
field at the front of the Serrania in the north, recently modeling of the maturation and migration of hydrocar-
extended into the Maturin basin by the discovery of the bons, to construct an integrated geologic model that
giant El Furrial field in the frontal buried overthrusts of describes the petroleum geology of the area.

Figure 1—Tectonic setting of the Eastern Venezuelan basin and locations of the major oil fields. Note the two main
petroleum provinces (Faja in the south and frontal units of the Serrania in the north).
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 743

the South American plate was diachronous with a


progressive west to east displacement (Stephan et
al., 1985).

In the central part of the Eastern Venezuelan basin


(Figure 1), only the two last episodes of this geodynamic
evolution can be recognized in wells and outcrops. The
first two are interpreted from seismic data in the
Trinidad subbasin to the east (Persad, 1978) and in the
Guarico subbasin to the west (Figure 1).

Prerift Megasequence
The prerift megasequence that developed during the
Paleozoic is identified on seismic profiles (Figures 2, 3). It
is associated with the Hato Viejo and Carrizal formations
of the Guarico subbasin which were deposited in coastal
to neritic marine environments. The sequences encoun-
tered by drilling are composed mainly of fine- to coarse-
grained sandstones, which are slightly calcareous and
intercalated with conglomerates and green shales. Stover
(1967) has dated the Carrizal Formation as Late
Devonian–Early Carboniferous. This sequence reaches a
thickness of about 1.5–2.0 sec (two-way time), or about
3000–5000 m.

Rift Megasequence
The rift megasequence developed during Late
Jurassic–earliest Cretaceous time and has been described
as the La Quinta Formation west of the study area in the
Espino graben (Hedberg, 1950; Bartok, 1993) (Figure 1).
This formation, which was deposited in a continental
environment, is composed mainly of red shales with
Figure 2—Map of study area showing locations of the
basaltic sills. This megasequence is seismically recog-
seismic profiles and structural sections. The study area nized in half-grabens (Figure 3) and reaches a thickness
includes 5000 km of seismic reflection lines and more than of 0.6 sec, or about 3600 m.
200 wells.
Passive Margin Megasequence
GEODYNAMIC EVOLUTION AND The passive margin megasequence covers the Creta-
STRATIGRAPHIC MODEL ceous–Paleogene and is characterized by three principal
transgressive phases that developed from north to south
The geodynamic evolution of the Eastern Venezuelan and culminated during Turonian, Paleocene–early
basin can be divided into four major episodes (Eva et al., Eocene, and Oligocene time, respectively (Figure 3).
1989): These events match the general eustatic sea level chart
(Vail et al., 1977). Although this megasequence is clear on
• A prerift phase in Paleozoic time seismic lines (Figure 3), the base has not yet been recog-
• A rifting and drifting phase during Jurassic and nized either in wells or in outcrop.
earliest Cretaceous time, characterized by The initial transgressive phase commenced with
grabens, the creation of oceanic crust in the deposition of the basal sandstone of the Barranquin
Tethyan-Caribbean domain, and a regional Formation (Van der Osten, 1957). The maximum trans-
breakup unconformity (Eva et al., 1989; Bartok, gressive advance is marked by the deposition of platform
1993) carbonates that are diachronous in a north-south
• A passive margin period during the direction within the basin and clearly seen on seismic
Cretaceous–Paleogene lines (Figure 3). This transgression is defined as the inter-
• A final phase during oblique collision in the mediate Cretaceous sequence. In the El Pilar area to the
Neogene and Quaternary that resulted in north (Figure 1), these limestones correspond to the
formation of the Serrania del Interior and transfor- upper level of the Barranquin Formation of Barremian
mation of the passive margin basin into a foreland age and, in the southernmost area of outcrops, to the
basin. This collision of the Carribean plate with Albian El Cantil Formation (Figure 4). In the El Furrial
744 Parnaud et al.

Figure 3—Seismic profile and structural interpretation of the Jurassic–Cretaceous sequence. See Figure 2 for location.

field, these limestones are of Aptian–Turonian age and During the Paleocene–Eocene, the next transgression
belong to the El Cantil, Querecual, and San Antonio followed the Maastrichtian regression; it is represented
formations. In the southern part of the basin (Mata area), by the San Juan Formation. In outcrops, the San Juan
these strata belong to the Tigre Formation and indicate Formation contains regressive submarine fans facies (Di
the maximum advance of the transgression during the Croce, 1989) and in the El Furrial area, deltaic-estuarine
Turonian (Figure 4). facies. This new facies encountered in one of the wells of
In the El Furrial area, another intermediate facies has the study area is laterally equivalent to the San Juan
been drilled between the Tigre and San Antonio forma- Formation and is referred to as the Musipan Formation
tions. It is characterized by sandstones rich in organic of Maastrichtian age (Figure 4).
matter, fossiliferous dolomitic limestones (with echino- The final transgression that developed during the
derm fragments), and the presence of glauconite. Oligocene commenced with deposition of the basal
Partially silicified phosphatic layers associated with sandstone of the Merecure Formation. This sandstone is
dolomitic limestones are common. the main reservoir of the El Furrial area (Figure 5). The
The main source rocks were deposited during this Merecure Formation was deposited in a continental envi-
transgressive phase (Talukdar et al., 1988; Gallango et al., ronment in the southern part of the basin (Cerro Negro
1992; Alberdi and Lafargue, 1993). They are organic-rich area) and in an inner shelf environment within the
marine mudstones deposited in a bathyal environment northern part (outcrops of the Serrania del Interior). It is
between the Pirital high and the El Pilar fault and a composed mainly of alternations of fine- to coarse-
mixed organic facies that occurs in a platform environ- grained sandstones and shales, the source of the
ment to the south between the Pirital high and the defor- sediments being the Guyana shield to the south. The
mation front. These source rocks correspond to the Merecure Formation changes facies toward the north; the
Querecual and San Antonio formations of Cenoman- sandstone shales out into the Areo and Carapita forma-
ian–Campanian age and are coeval with the deposition tions, which are attributed to an outer shelf environment
of platform carbonates to the south (Figure 3). (Stainforth, 1971).
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 745

Figure 4—North-south stratigraphic correlation chart along the regional transect.

Oblique Collision Megasequence main seal for the Oligocene–Miocene reservoirs (Freites
and Oficina formations) (Figure 5) and is also regarded
The passive margin megasequence ended during the as a potential source rock with mixed organic matter
Oligocene when it was terminated by the collision of the types (continental and marine). However, the foredeep
Carribbean plate against the South American plate and was mainly filled during the Pliocene–Pleistocene with
the basin changed into a foreland basin. The oblique the younger La Pica and Las Piedras deposits, represen-
collision migrated progressively eastward during the late tative of shallow marine to continental environments
Oligocene to earliest Miocene, dividing the foreland (Figure 5). The maximum cumulative thickness of the
basin into three areas (Figure 1): (1) a southern area (from oblique collision megasequence is about 6000 m.
Cerro Negro to Oritupano) corresponding to the Near the deformation front and in the allochthon, the
platform zone, (2) a central zone (from Acema-Casma to synorogenic deposits contain numerous unconformities
Pirital) corresponding to the foredeep, and (3) a northern that help to separate the various formations (Figure 5)
area (north of the Pirital fault) corresponding to the over- and date the deformation:
thrust area.
• The lower Miocene Naricual Formation (Socas,
Diachronism of the thrust belt load applied to the
1990) records the initiation of uplift of the Serrania
northern border of the South American continent led to
del Interior in the north and a coeval southward
an eastward migration of the foredeep (Stephan et al.,
propagating deltaic system.
1985). In the west, the latest passive margin megase-
• The Chapapotal Member is related to the Pirital
quence sedimentary rocks are Paleocene, thus indicating
thrust of middle Miocene age.
early tectonic events in that area. In eastern Venezuela,
• The Morichito Formation corresponds to the conti-
however, no tectonic activity has been recognized in the
nental infill of a piggy-back basin transported by
rock record prior to the late Eocene. However, the
the Pirital thrust (Roure et al., 1994).
foredeep also migrated from north to south, which was
• Turbiditic sedimentary deposits of the lower
related to stacking of the allochthonous units caused by
Carapita Formation have a possible northwest to
collision of the plates (Rossi, 1985; Potié, 1989).
southeast transport direction.
The foredeep, located south of the deformation front,
was isolated from the platform by the slope. In the Each of these formations, except for the Morichito, were
foredeep itself, the thick shale sequence of the Carapita deposited in bathyal environments in front of the thrusts
Formation was deposited first. This formation is the that generated them.
746 Parnaud et al.

Figure 5—Conceptual cross sections showing the deposition of Oligocene and early Miocene sediments. (A) Regional
model; (B) local model.

STRUCTURAL CROSS SECTIONS AND American plate boundary) to the basin axis corresponds
to a south-vergent system. Its compressive structural
MAJOR PETROLEUM TRAPS style is characterized by the following:
The analysis of more than 5000 km of seismic reflec- • Thin-skinned tectonic units consist of Cretaceous–-
tion lines has been supplemented with surface geologic Tertiary deposits that are detached from their
data, computer-assisted techniques to restore structural substratum along a major décollement at the
cross sections (Roure et al., 1994), and integration of Jurassic–Cretaceous interface (evaporites or coal
gravity and magnetic modeling to construct a coherent measures located at an average depth of 10–15 km)
structural model (Figure 6). (Figure 6) (Roure et al., 1994). Shallower detach-
ments occur elsewhere in the tectonic pile, espe-
Structural Styles cially at the boundary between the indurated
Two different tectonic provinces have been recog- Mesozoic and Paleogene platform deposits and the
nized. The autochthonous province extends from the poorly lithified Neogene Carapita Formation.
basin axis to the Orinoco River. It is extensional and There is probably another décollement level within
characterized by normal faults trending N 60º–70º E the Paleozoic–Jurassic succession.
(Figures 7, 8) and transcurrent faults that are younger in • The deformation front of the allochthonous
the southern part than in the north. Strike-slip faults province locally corresponds to a triangle zone,
trending N 70º W are also present in the northern part of with progressive wedging of the foredeep
the province, affecting both Cretaceous and Paleogene Carapita Formation by the Mesozoic and
deposits. Northward-dipping gravity collapse faults Paleogene sequences in the El Furrial ramp
affect the Miocene–Pliocene sedimentary rocks. anticline (Figure 9). Other fishtail or triangle struc-
The allochthonous province that extends from the El tures occur behind the deformation front, that is,
Pilar fault (a major east-trending, dextral strike-slip fault along the Pirital high.
representing the surface trace of the Caribbean–South • Kilometer-scale cylindrical folds and associated
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 747

Figure 6—Structural cross sections across the Maturin basin from the Serraniato to the Faja. See Figure 2 for location. (A)
Present stage. (B) Initial stage. (Modified after Roure et al., 1994; Passalacqua et al, 1995.)

Figure 7—Seismic profile showing normal faulting in the foreland platform area. See Figure 2 for location.

thrusts trending N 60º–70º E are mainly ramp anti- (Roure et al., 1994).
clines, such as in the El Furrial area (Figures 6, 10). • Dextral transcurrent faults with a N 50º–60º W
• The thrust sequence is mainly a piggyback direction (Urica, San Francisco, and Los Bajos
sequence, with local out-of-sequence thrusting or faults) (Figures 1, 9) separate the Serrania del
fault reactivation in the inner parts of the tectonic Interior into areas of diachronous structural
wedge (Pirital and Manresa faults) (Figure 6) evolution (Munro and Smith, 1984).
748 Parnaud et al.

Figure 8—Seismic profile showing traps in the Faja. See Figure 2 for location.

An intermediate zone showing compressive and However, the Urica fault only represents a tear fault that
extensional structural styles can be interpreted between borders the allochthon in the west. This fault does not cut
the allochthonous and autochthonous provinces. through the basement, with the initial border fault of the
Mesozoic inverted basin being located to the east (Roure
Balanced Structural Cross Sections et al., 1994).

Depth-converted line drawings compiled from the Geodynamic Evolution and Diachronous
seismic profiles and additional well and outcrop data Tectonics
have been used to build preliminary geologic cross
sections from the El Pilar fault in the north to the Orinoco Integration of the focal mechanisms of earthquakes
River in the south (Figure 1). They were then balanced with gravity and magnetic modeling of the crustal archi-
with the Locace software, an example of which is shown tecture of the belt (see Passalacqua et al., 1995) indicates a
in Figure 6. Details of these restorations are given minimum amount of type A subduction of the South
elsewhere (Passalacqua et al., 1995; Roure et al., 1994). American lithosphere beneath the Caribbean plate. The
These balanced cross sections show a shortening amounts of subduction are too limited to develop associ-
discrepancy between the passive margin sedimentary ated volcanism. The regional geodynamics are effectively
rocks (about 80 km) and between the pre-Cretaceous dominated by oblique collision (transpression) between
deposits and the upper crust (about 45 km). This differ- the Caribbean and South American plates. From Eocene
ence implies that the upper crust is also involved in the time until the present, the Caribbean plate and associated
shortening process, with a deep detachment either structural stresses have migrated eastward. This
between the brittle (upper) and ductile (lower) crust or at migration caused diachronism in tectonics and sedimen-
the crust–mantle interface. tation along the entire northern border of Venezuela, as
Moreover, the pattern of compressive deformation well as strain partitioning across the plate boundary
was probably influenced by the history of the crust (Stephan et al., 1985; Passalacqua et al., 1995).
during the rift phases (inversion of the previous normal This west to east migration and partitioning of the
faults). This inversion is best imaged along eastward- compressive stress is shown by the growth of major
trending seismic profiles that cross the Urica fault zone structural elements:
(Figure 9), where thick basinal strata of a Mesozoic
subbasin (El Furrial and Pirital units) are thrust over • Progressive eastward displacement of the
thinner and still autochthonous strata of the foreland. Caribbean plate along the El Pilar fault
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 749

Figure 9—Seismic profile showing the Urica transfer fault. See Figure 2 for location.

• Progressive emplacement of the various compart- Reservoir Quality


ments of the Serrania allochthon along transverse
transcurrent faults (Urica, San Francisco, and Los The results of the reservoir study show that porosity
Bajos faults) values are generally high (up to 35%) for shallow reser-
• Mainly north-south shortening along the N 60º– voirs such as Las Piedras, Oficina, and Merecure in the
70º E trending frontal thrusts (El Furrial and Pirital) Orinoco heavy oil belt and are reduced by compaction
due to depth of burial. A general compaction law
(Gallango et al., 1992) for the Las Piedras, Freites,
Oficina, and Merecure reservoirs south of the deforma-
RESERVOIR AND HYDRODYNAMIC tion front is
MODELS
South of the Serrania del Interior outcrops, ten PHI= e 3.612–depth/12708
different reservoirs have been recognized and studied
within the Chacopata-Uverito subsurface transect. The where PHI is the effective porosity in percent.
studied Lower Creataceous–Pleistocene reservoirs
include the San Juan and other undifferentiated Creta- In addition to these primary obliteration processes, reor-
ceous; the Paleogene Merecure and Caratas Formations; ganization of the porous media also occurs, especially
and the Neogene Las Piedras, La Pica, Morichito, Chapa- through diagenetic processes such as dissolution and
potal, Freites, and Oficina sandstones (Figures 3, 5, 11). fracturing. These processes have been recognized in the
By integrating pressure measurements, geochemical Aguasay deep zones and in the El Furrial slab, where up
analyses, and new structural data, a hydrodynamic to 60% of the encountered porosity is of secondary origin
model is proposed for the Eastern Venezuelan basin. (Aguado et al., 1993).
750 Parnaud et al.

Figure 10—Seismic profile showing the El Furrial structure. See Figure 2 for location.

From an exploration prospective, three main areas are shallow for the surface aquifer or deeper for the
recognized (Figure 1, 11, 12): Oligocene–Miocene aquifer. In contrast, the main
discharge was restricted to the Orinoco River banks, with
• The platform area from Aguasay-Casma to Cerro a secondary discharge in the delta zone for the surface
Negro, where the objective is the Oligocene–- aquifer. A deep north to south drainage thus existed
Miocene reservoir with intergranular primary within most of the subbasin area, while in the southern
porosity and western parts, the flow direction at shallow depths
• The El Tejero-El Furrial-Orocual thrust zones, was from west to east. Because vertical communication
where the objective is the Merecure sandstone with exists among the different aquifers in the Orinoco heavy
secondary and primary porosity oil belt, this direction is also preserved there, so that the
• The Pirital high zone, where the objectives are the drainage within the deeper Oligocene–Miocene reservoir
Las Piedras, Morichito, and Merecure sandstones, is from west to east.
with primary porosity, and the Cretaceous lime- Due to later thrusting, the deep aquifer is no longer
stones with secondary fracture porosity continuous; it has been subdivided so that the meteoric
refill area of the Serrania is now bypassed. The Las
Hydrodynamic Model Piedras transit paths were not modified, but the Oligo-
cene–Miocene drainage pattern was totally reorganized.
Based on pressure data and formation water composi- In the eastern part (Oritupano area) (Figure 12), a
tion distributions, combined with the integration of new connate water refill area is recognized that was formerly
structural data, a hydrodynamic model is proposed for hidden by the large flow of meteoric water. This connate
the subbasin (Figure 12). This model involves two phases water is the product of compaction (dewatering) of the
separated by the last thrust emplacement of the El Furrial thick foredeep deposits. The El Furrial–Orocual thrust
and Pirital structures. unit is isolated and characterized by mixed formation
Prior to recent events, two major topographic highs waters and high pressures, reaching 75–80% of the litho-
may have represented the necessary recharge area for static load. Along the Orinoco heavy oil belt, the pattern
meteoric refill of the Las Piedras and Oligocene–Miocene was not changed; meteoric waters still migrated in each
aquifers: the Serrania del Interior to the north and the aquifer from the Mesa highs to reach their exsurgence
paleohighs to the west. The transit zones were either along the Orinoco River banks.
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 751

Figure 11—Conceptual cross section showing the major petroleum plays of the Eastern Venezuelan basin.

From the study of the crude oil distribution, we


recognize a migration trend. During the first phase, the
oil probably migrated rapidly toward the Orinoco fields
due to the large freshwater flow arising from the uplifted
ranges to the north. The “tar” and heavy biodegrated oils
encountered everywhere in the traps of the basin are
regarded as evidence for such a process. During the
second phase, the lack of a strong hydrodynamic vector
confined the migration of the newly generated oils to the
Onado-Aguasay and northern areas. Thus, oil migration
was no longer in a north-south direction but in an east-
west direction. To the west and southwest, the newly
generated oils are presently undergoing biodegradation
due to the meteoric water percolating from the Mesa
highs.

OIL KITCHENS AND MIGRATION PATHS


Extensive geochemical sampling has been undertaken
both in the Serrania (outcrops) and in the wells of the
Maturin basin. Results of Rock-Eval pyrolysis (Espitalié
et al., 1985) and complementary analyses give a good
estimate of the distribution and quality of the potential
source rocks in the area (Talukdar et al., 1988; Alberdi
and Lafargue, 1993) (Figure 13). One-dimensional simu-
lations of the maturation of the organic-rich deposits
were undertaken using GENEX software, whereas two-
dimensional migration paths were tested with THEMIS
software. A detailed description of the numerical
modeling is given elsewhere (Gallango and Parnaud,
1995).

Source Rocks
Three distinct source rocks have been identified. The
principal source rock includes parts of the Querecual and
San Antonio formations (Figures 4, 13). The organic
matter has a marine origin, with an average TOC content
of 2–6 wt. % and a petroleum potential higher than 5 mg
HC/g rock. This source rock is widely distributed in the
Serrania del Interior. A secondary source rock lies within
the San Antonio and Querecual formations (Figures 4, Figure 12—Map showing the hydrodynamic system of the
13), but it has a mixed organic matter type, mainly conti- Maturin subbasin.
752 Parnaud et al.

Figure 13—Hydrogen index (HI) versus oxygen index (OI) and HI versus Tmax diagrams of the major source rocks. (Units are
mg HC/g TOC for HI and mg CO2/ g TOC for OI.)
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 753

nental, with an average initial TOC of 2 wt. % and a


petroleum potential of about 2 mg HC/g rock. It is
distributed mainly in the Pirital and El Furrial over-
thrusts and in the lateral equivalent of the Querecual
Formation in the autochthonous province to the south.
Finally, the least significant source rock is distributed in
scattered levels of the Carapita Formation in the
autochthonous part of the foredeep basin (Figures 5, 13).
It has a mixed organic matter type, mainly continental,
with an average original TOC of 2 wt. % and a petroleum
potential of 2–5 mg HC/g rock.

Hydrocarbon Kitchens
Numerical modeling has been applied to simulate the
maturation of the Cretaceous potential source rocks
(Gallango and Parnaud, 1995). The Carapita source rock,
which is essentially gas prone, has not been investigated
because of the lack of information. However, it is
probably the source for numerous biogenic gas shows
encountered in the foredeep. Despite its poor oil
potential (type III organic matter of continental origin), it
could have generated small quantities of oil. The
discovery of terrestrial oil in a recent well may relate to
the Carapita source rock, unless it is derived from the
terrestrial organic matter contained in parts of the
Querecual Formation.
Results from numerical simulations of the maturation
of the source rocks help to identify three major kitchens
(Figures 14, 15). The first is located in the autochthon and
El Furrial areas. It contributed to a reduced petroleum
generation potential from Miocene to present time. The
source rocks of this domain are presently in the gas Figure 14—Oil maturity distribution along the Chacopata-
window and have a high transformation ratio. Uverito transect. See Figure 2 for location.
The second source rock kitchen corresponds to the
Pirital slab. It has a moderate petroleum generation
potential and acquired its present maturity during the but eroded sedimentary sequences. In the Pirital area, oil
early Miocene (episode of maximum burial). Petroleum generation started very early, by the end of Cretaceous
was generated in this area from Paleocene to middle time. A later phase of oil generation was initiated during
Miocene time. The maturity index varies from south to the thrusting episode and provided more mature crudes,
north, with vitrinite reflectance values of 0.65–1.30% Ro but in limited quantities. Without considering possible
in the present overthrust zone and a transformation ratio gas generation, between 300 and 930 billion bbl of oil
of 15–95% in the area investigated. Petroleum generation have been generated in this area.
probably occurred during two distinct episodes in the In conclusion, the total amount of generated oil within
Pirital allochthon. During the first episode, prior to the study area is estimated to be between 420 and 1350
thrusting, an immature crude was probably generated billion bbl.
that migrated southward toward the Orinoco subbasin.
In the second episode, a minor amount of more mature Origin of Hydrocarbons and Trapping
crude was also generated. The total amount of petroleum
generated in the Pirital kitchen is estimated to be On the basis of the distribution of oil fields (Figure 1)
between 120 and 420 billion bbl, without taking into and the simulation of hydrocarbon generation, various
account gas generation (Parnaud et al., 1992b). migration paths and episodes of trapping are envisaged
The third kitchen relates to the Serrania del Interior (Figure 15). To the south in the Cerro Negro area (Figures
itself and is characterized by a high petroleum potential. 1, 15), the oil has a marine signature. It is clearly derived
This area is presently in an uplifted position, and its from the first generation of the marine source rocks of
maturity level does not fit with the present burial depths. the Querecual and San Antonio formations, which began
This high maturity level was reached prior to thrusting in early Paleocene time in the Serrania del Interior area.
during early Miocene time. It implies either a high geo- In the Acema-Casma central area (Figures 1, 15), there
thermal gradient at that time or the recent erosion of an is a mixture of mature and immature oils, all of which
overlying tectonic unit (tectonic overburial) or unknown were derived from a marine source. As in the Cerro Negro
754 Parnaud et al.

The gas that accumulated in the eastern part of the El


Furrial trend (El Corozo structure) is thermogenic and
associated with oil, although in the western part of the
trend (El Tejero–El Carito structures), the gas is associ-
ated with condensate. The gas encountered in the
Onado–Acema Casma fields is also thermogenic and
associated with oil. All the analyzed gas samples from
these fields show that they were generated principally
from a marine source rock, compatible with the San
Antonio and Querecual source proposed for the oils.
In the Pirital area (Figures 1, 6, 15), the occurrence of
an immature to mature crude oil of marine origin is
probably related to the last episode of oil generation in
the Querecual and San Antonio formations. This episode
took place after the latest thrusting and subsequent
isolation of the Pirital unit from other parts of the basin.
Newly generated oils migrated through faults and
fractures toward the shallower reservoirs of the La Pica,
Morichito, and Las Piedras reservoirs. Secondary
migration or remigration of early generated oils toward
the surface may reflect recent reactivation of the Pirital
overthrust fault as part of an out-of-sequence thrust
(Roure et al., 1994) and represents an alternate solution
for the oils trapped in the Morichito piggyback basin.

CONCLUSIONS
This integrated geologic model of the Eastern
Venezuelan basin provides crucial data on the sedimen-
tary fill, internal structural architecture, and evolution of
the basin. Four major episodes of basin evolution are
recognized: prerift, rift, passive margin, and foreland
basin. The change from passive margin to foreland basin
evolution is dated as latest Oligocene or early Miocene. It
is essential to distinguish the Merecure Formation in the
El Furrial area, which is attributed to the Oligocene
passive margin, from the Naricual Formation whose
Figure 15—Tectonic evolution of northern Venezuela. Serrania outcrops belong to the early Miocene foreland
Conceptual cross sections showing the locations of oil basin. The outcrop equivalent of the Merecure Formation
kitchens and changes in migration paths during the geo- is the Oligocene Los Jabillos Formation. A new correla-
dynamic evolution. (A) Passive margin stage: source rock tion of the Cretaceous sequences is proposed for the El
deposition. (B) Passive margin stage: early maturation, Furrial area. The Musipan Formation is believed to be a
expulsion, and migration. (C) Type B subduction: long southern equivalent of the San Juan Formation, whereas
distance migration to the Faja petrolifera del Orinoco. the El Furrial Formation is a lateral equivalent of the San
(D) Type A subduction: late and short distance migration to
the El Furrial and Pirital traps; present-day kitchen.
Antonio Formation (to the north) and the Tigre
Formation (to the south).
The cross section balancing has constrained the burial
area, these crude oils correspond to the first generation history of the potential source rocks, enabling us to
from the Querecual and San Antonio formations, starting simulate the petroleum generation. Major results of this
in the Paleocene in the area of present-day deformation. study address the petroleum potential of the basin. The
In the Central Onado and northern Furrial areas differences observed among the crude oils are due to the
(frontal thrust units) (Figures 1, 6, 15), the occurrence of a effects of biodegradation, maturity variations, and
very mature crude oil of marine origin is related to a natural mixing processes. Most of the oils were derived
second episode of generation that was associated with from a marine source: the San Antonio and Querecual
the rapid tectonic overburial of the Querecual and San Cretaceous formations.
Antonio formations. During this episode, part of the The source rock for the terrestrial crude oil has not
generated crudes probably still migrated southward and been identified. The Querecual, San Antonio, or Carapita
reached the Onado traps, although most of the oil could formations could have generated this crude. Distinct
migrate no farther than the frontal structures (e.g., El episodes of maturation, expulsion, and migration of
Furrial). hydrocarbons are recognized. Early maturation mainly
Petroleum Geology of the Central Part of the Eastern Venezuelan Basin 755

resulted from sedimentary overburial, with the Colombia, Ecuador y Peru: Asociación Colombiana de
generated oils having migrated long distances from the Geología y Geofísica de Petróleo, Bogota, 45 p.
present Serrania area toward the flexural bulge (Orinoco Gallango, O., and F. Parnaud, 1995, Two-dimensional
fields). In contrast, late stage maturation was related to computer modeling of oil generation and migration in a
transect of the Eastern Venezuelan basin, in A. J. Tankard,
tectonic overburial, with the resulting crudes having
R. Suarez, and H. J. Welsink, Petroleum basins of South
migrated short distances to fill the frontal traps of the America: AAPG Memoir 62, this volume.
allochthon (El Furrial field). Gallango, O., M. Escandon, M. Alberdi, F. Parnaud, and J.-C.
Pascual, 1992, Hydrodynamism, crude oil distribution, and
geochemistry of the stratigraphic column in a transect of
the eastern Venezuela basin (abs.): GSA, Abstracts with
Programs, p. A214.
Acknowledgments The authors thank Intevep, S.A., for Gonzalez de Juana, C., J. Arozena, and X. Picard-Cadillat,
permission to publish this paper and the industry sponsors 1980, Geología de Venezuela y de sus cuencas petrolíferas:
Corpoven, S.A., Lagoven, S.A., and Maraven, S.A., opera- Caracas, Foninves ediciones, v. 2, 1051 p.
tional affiliates of PDVSA, for the financial support of the Hedberg, J. D., 1950, Geology of the eastern Venezuela basin
project and for providinge the subsurface data required for the (Anzoategui-Monagas-Sucre-eastern Guarico portion):
GSA Bulletin, v. 61, p. 1173–1216.
study. This paper has also been greatly improved thanks to the
Munro, S. E., and F. D. Smith, Jr., 1984. The Urica fault zone,
comments of M. Cooper, M. Daly, A. J. Tankard, and an northeastern Venezuela: GSA Memoir, v. 162, p. 213–215.
anonymous reviewer. Parnaud, F., I. Truskowski, Y. Gou, J. Di Croce, and J.-C.
Pascual, 1992a, Stratigraphic model of the Chacopata-
Uverito transect (eastern Venezuelan basin) (abs.): GSA,
Abstracts with Programs, p. A360.
REFERENCES CITED Parnaud, F., I. Truskowski, Y. Gou, O. Gallango, J.-C. Pascual,
J. Di Croce, F. Roure, and H. Passalacqua, 1992b, Inte-
grated geological model in a transect of the eastern
Aguado, B., L. Ramirez, and M. Quintero, 1993, Sedimen- Venezuelan basin (abs.): GSA, Abstracts with Programs,
tology of the Boquerón field, Eastern Venezuela basin p. A309.
(abs.): AAPG and Society of Venezuelan Geologists, Passalacqua, H., F. Fernandez, Y. Gou and F. Roure, 1995,
Caracas International Congress and Exhibition, Abstract Deep architecture and strain partitioning in the eastern
volume, p. 33. Venezuelan ranges, in A. J. Tankard, R. Suarez, and H. J.
Alberdi, M., and E. Lafargue, 1993, Vertical variations of Welsink, Petroleum basins of South America: AAPG
organic matter content in Guayuta Group (Upper Creta- Memoir 62, this volume.
ceous), interior mountain belt, eastern Venezuela: Organic Persad, K. M., 1978, Hydrocarbon potential of the Trinidad
Geochemistry, v. 20, p. 425–436. area: Geologie en Mijnbouw, v. 57, p. 277–285.
Arnstein, R., E. Cabrera, F. Russomanno, and H. Sanchez, Potié, G., 1989, La Serrania del Interior oriental sur le transect
1985, Revisión estratigráfica de la cuenca de Venezuela Cumana-Urica et le bassin de Maturin (Venezuela): Ph.D.
oriental: VI Conferencia Geológica Venezolana, Memoria I, dissertation, Université Bretagne Occidentale, Brest, 240 p.
p. 41–69. Rossi, T., J. F. Stephan, R. Blanchet, and G. Hernandez, 1987,
Bartok, P., 1993, Prebreakup geology of the Gulf of Etude géologique de la Serrania del Interior oriental sur le
Mexico–Caribbean: its relation to Triassic and Jurassic rift transect Cariaco-Maturin: Revue de l’Institut Français du
systems of the region: Tectonics, v. 12, p. 441–459. Pétrole, Paris, v. 42, no. 1, p. 3–30.
Carnevali, J. O., 1988a, El Furrial oil field, northeastern Rossi, T., 1985, Contribution à l’étude géologique de la
Venezuela: first giant in foreland fold and thrust belts of frontière sud-est de la plaque caraïbe: la Serrania del
western hemisphere (abs.): AAPG Bulletin, v. 72, p. 168. Interior oriental (Venezuela) sur le transect Cariaco-
Carnevali, J. O., 1988b, Venezuela nor-oriental: exploración en Maturin: Ph.D. dissertation, Université Bretagne Occiden-
frente de montaña, in A. G. Bellizzia, A. C. Escoffery, and I. tale, Brest, 338 p.
Bass, eds., III Simposio Bolivariano: Sociedad Venezolana Roure, F., J. O. Carnevali, Y. Gou, and T. Subieta, 1994,
de Geología., Caracas, v. 1, p. 69–89. Geometry and kinematics of the north Monagas thrust belt
Chiock, M., 1985, Cretáceo y Paleógeno en el subsuelo de (Venezuela): Marine and Petroleum Geology, v. 11,
Monagas: VI Conferencia Geológica Venezolana, Caracas, p. 347–362.
v. 1, p. 350–383. Socas, M., 1990, Estudio sedimentologico de la Formación
Di Croce, J., 1989, Análisis sedimentológico de la Formación Naricual, Estado Anzoategui: B.Sc thesis, Universidad
San Juan en la Cuenca Oriental de Venezuela (Estados Central de Venezuela, Caracas, 421 p.
Anzotegui y Monagas): Master’s Thesis, Universidad Stainforth, R. M., 1971, La Formación Carapita de Venezuela
Central de Venezuela, Caracas, 205 p. oriental: IV Congreso Geológico Venezolano, Caracas,
Espitalié, J., G. Deroo, and F. Marquis, 1985, La pyrolyse Boletin Geologico, I (V), p. 433–463.
Rock-Eval et ses applications (part II): Revue de l’Institut Stephan, J. F., R. Blanchet, and B. Mercire de Lepinay, 1985,
Français du Pétrole, v. 40, p. 755–784. Les festons nord et sud caraïbes (Hispaniola–Porto Rico;
Eva, A., K. Burke, P. Mann, and G. Wadge, 1989, Four-phase Panama et Colombie-Vénézuela): des pseudo-subductions
tectonostratigraphic development of the southern induites par le raccourcissement est-ouest du bâti conti-
Caribbean: Marine and Petroleum Geology, v. 6, p. 9–21. nental péri-caraïbe, in A. Mascle, ed., Géodynamique des
Fiorillo, G., 1982, Exploración y evaluación de la Faja Caraïbes: Paris, Edition Technip, p. 35–52.
Petrolífera del Orinoco, in Symposium Exploración Stover, L. E., 1967, Palynological dating of the Carrizal
Petrolera en las Cuencas Subandinas de Venezuela, Formation of eastern Venezuela: Asociación Venezolana
756 Parnaud et al.

de Geólogos Mineros y Petroleros, Boletin, v. 10, Authors’ Mailing Addresses


p. 288–301.
Sulek, J. A., 1961, Miocene correlations in the Maturin François Parnaud
subbasin: Asociación Venezolana de Geólogos Mineros y Yves Gou
Petroleros, v. 4, p. 131–139. Jean-Claude Pascual
Talukdar, S., O. Gallango, and A. Ruggiero, 1988, Generation Beicip-Franlab
and migration of oil in the Maturin subbasin, eastern B.P. 213
Venezuelan basin, in L. Matavelli and L. Novelli, eds., 92502 Rueil-Malmaison
Advances in organic geochemistry: Organic Geochemistry, France
v. 13, p. 537–547.
Vail, P. R., R. M. Mitchum, Jr., and S. Thompson, 1977, Seismic Irene Truskowski
stratigraphy and global changes of sea level, part 4: global
Oswaldo Gallango
cycles of relative changes of sea level, in C. E. Payton, ed.,
Seismic stratigraphy—application to hydrocarbon explor- Herminio Passalacqua
ation: AAPG Memoir 26, p. 83–97. Intevep, S.A.
Van der Osten, E., 1957, Lower Cretaceous Barranquin Apartado Postal 76343
Formation of northwestern Venezuela: AAPG Bulletin, v. Caracas 1070A
41, p. 679–708. Venezuela

F. Roure
IFP
B.P. 311
92506 Rueil-Malmaison
France
Cenozoic Sedimentation and Tectonics of the
Southwestern Caribbean Pull-Apart Basin,
Venezuela and Colombia

C. E. Macellari
Shell Oil
Houston, Texas, U.S.A.

Abstract

T he complex Caribbean–South American plate boundary records a Late Cretaceous–Eocene phase of


terrane collision followed by Eocene–Recent right-lateral displacement. This study analyses the strati-
graphy deposited during the latter stage in a series of episutural pull-apart basins. On the basis of regional
seismic, well control, field work, and published information, this succession is divided into four unconformity-
bounded depositional sequences. These cycles are upper Eocene–lower Oligocene, upper Oligocene–lower
Miocene, middle–upper Miocene, and Pliocene–Recent. The remnants of upper Eocene and lower Oligocene
rocks are mostly restricted to narrow, fault-controlled northwest-southeast depocenters. During the late
Oligocene–late Miocene, sedimentation was still fault controlled but more widespread. Sedimentation rates
along these growth faults were extremely high (up to 350 m/m.y.), but generally decreased through time.
During the Eocene, the axis of maximum subsidence was located in the western part of the Golfo de
Venezuela, at the contact between autochthonous and allochthonous units emplaced during a prior collisional
event. During the Oligocene and Miocene, the axis of maximum subsidence was located farther east, in the
Urumaco trough and east of La Vela Bay. At the same time, an ENE-WSW oriented depocenter began to
develop in the Falcón basin in response to loading by a northward-advancing thrust front. Finally, during the
Pliocene, the largest subsidence rates occurred south of the Golfo Triste, in a pull-apart area created by the
right-lateral displacement of the Boconó and Oca faults.

Resumen

E l complejo contacto de las placas del Caribe y Sudamérica registra una fase cretácica tardía a eocena de
colisión de terrenos, que fue seguida por desplazamientos laterales derechos durante el Eoceno al
Reciente. Este estudio analiza la estratigrafía depositada durante esta última etapa en una serie de cuencas
“pull-apart” episuturales. Basado en el estudio de sísmica regional, pozos, trabajo de campo, y análisis de liter-
atura, esta sucesión se divide en cuatro secuencias deposicionales limitadas por discordancias. Estos ciclos son:
Eoceno superior–Oligoceno inferior; Oligoceno superior–Mioceno inferior; Mioceno medio–superior, y
Plioceno–Reciente. Los remanentes de depositación del Eoceno tardío y del Oligoceno temprano están mayor-
mente restringidos a depocentros angostos orientados noroeste-sudeste, que fueron controlados por fallas.
Durante el Oligoceno tardío al Mioceno tardío, la sedimentación aún estuvo controlada por fallas, pero fue mas
distribuida arealmente. Los ritmos de sedimentación en estas fallas de crecimiento fueron extremadamente
altos (hasta 350 m/m.a.), pero decrecieron en general a travéz del tiempo.
Durante el Eoceno, el eje de máxima subsidencia estuvo ubicado en la parte occidental del Golfo de
Venezuela, en el contacto entre unidades autóctonas y alóctonas emplazadas durante un evento collisional
previo. Durante el Oligoceno y el Mioceno, el eje de subsidencia máxima estuvo ubicado mas hacia el este, en
el surco del Urumaco, y al este de la Ensenada de La Vela. Al mismo tiempo un depocentro orientado este
noreste-oeste sudoeste comenzó a desarrollarse en la cuenca de Falcón en respuesta flexural a la carga
producida por un frente de corrimiento que avanzaba desde el sur. Finalmente, durante el Plioceno el ritmo
mas alto de sedimentación se localizó al sur del Golfo Triste, en una cuenca “pull-apart” creada por el
desplazamiento lateral-derecho de las fallas de Boconó y Oca.

Macellari, C. E., 1995, Cenozoic sedimentation and tectonics of the southwestern 757
Caribbean pull-apart basin, Venezuela and Colombia, in A. J. Tankard, R. Suárez S., and
H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 757–780.
758 Macellari

Figure 1—Tectonic and basin map of northwestern South America and location of study area. (Modified from Stephan, 1985.)

INTRODUCTION TECTONIC PROVINCES


Most of the Caribbean–South American plate The southern Caribbean margin is formed by a suite
boundary is represented by a regional right-lateral of tectonic terranes that include the following, from north
transpressive contact. A series of Cenozoic basins to south (Figure 1):
developed along this contact as a response to plate
interactions, which included an early collision and a South Caribbean Deformed Belt—The south Carib-
subsequent pull-apart stage (Muessig, 1984; Pindell, bean deformed belt (Case et al., 1984) and its eastern
1991). This complex geologic history is recorded in the continuation, the Curaçao Ridge, are Neogene accre-
sedimentary succession. tionary complexes that formed in a compressional
The aim of the present study is to describe and regime as the Caribbean borderlands were thrust
interpret the Cenozoic depositional history of the north- northward over the Caribbean basin (Silver et al., 1975;
western South American margin to better understand Talwani et al., 1977; Ladd et al.,1984).
the evolution of the area (Figure 1). To achieve this, the Los Roques Basin—The Los Roques basin, located
stratigraphy was simplified in terms of unconformity- between the Curaçao Ridge and the Leeward Antilles,
bounded depositional sequences or cycles. These cycles contains more than 2000 m of Tertiary pelagic and
were identified on the basis of prominent unconformities possibly turbiditic strata overlying oceanic crust (Silver et
in the seismic record, regional paleontologically dated al., 1975; Case et al., 1984). The Los Roques basin was
hiatuses, marked changes in depositional environments, interpreted by Silver et al. (1975) as a secondary feature
widespread distribution of characteristic lithologic types, associated with uplift north of the Curaçao Ridge.
and changes in the areal distribution of stratigraphic Leeward Antilles Terrane—The Leeward Antilles (or
units, including variations in coastal onlap. The database Netherland Antilles) terrane is a series of uplifted blocks
for this study includes published literature, several trending eastward off the northern coast of Venezuela.
regional seismic lines, information from 25 wells, and These blocks are exposed in the Aruba, Curaçao, Bonaire,
field observations in the Falcón-Paraguaná area. Las Aves, and Los Roques islands. The basement rocks
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 759

are middle(?) Cretaceous tholeiitic pillow basalt and of these is the Oca fault, which has a documented
diabase of possible oceanic crust origin. Upper Creta- displacement of at least 15–20 km (Feo-Codecido, 1972;
ceous rocks interpreted as a primitive magmatic arc are Tschanz et al., 1974).
superimposed (Beets, 1972; Silver et al., 1975). Uplift and
erosion during the Paleogene resulted in deposition of a
conglomerate that is exposed on central Bonaire (Pijpers,
1933; Beets et al., 1977). This conglomerate includes clasts STRATIGRAPHY
of chert, volcanic rocks, and serpentinite, along with
fragments of gneiss, marble, and various schists, which The present study concentrates on the Eocene and
may indicate a continental provenance (Pijpers, 1933). younger strata that were deposited around the South
Bonaire Basin—The elongated Bonaire basin extends American–Caribbean plate margin. To simplify and
south of the islands of Curaçao and La Orchila and better understand equivalent events throughout the area,
contains more than 4000 m of Cenozoic deposits (Case et the complex Cenozoic stratigraphy was analyzed in
al., 1984). This basin is an offshore extension of the Falcón terms of unconformity-bounded depositional sequences.
basin (Silver et al., 1975; Case et al., 1984), but it is Four major cycles (D, C, B, and A) extending from the
underlain by a transitional to oceanic crust (Edgar et al., upper Eocene to the Pliocene were identified (Figure 2).
1971; Case et al., 1984). Because of the diversity of facies and terminology, these
Cordillera de la Costa—The Caribbean system in cycles are described independently for each geographic
north-central Venezuela is represented by the Cordillera area.
de la Costa, which is composed of a series of east-west-
trending tectonic belts. These consist of Mesozoic Guajira Peninsula
metasedimentary and metavolcanic rocks that were
emplaced south- or southeastward over Paleogene flysch Tertiary sedimentary rocks in the northern and
(Bellizzia, 1972; Maresch, 1974; Beck, 1978). eastern Guajira Peninsula area are preserved in three
Guajira Peninsula—The Guajira Peninsula has a core basins (Figure 3): the Cocinetas, Portete, and
of Cretaceous metamorphic rocks that are exposed in the Chichibacoa.
Serranias de Cocinas, Macuire, and Jarara and are
surrounded by relatively thin and mildly deformed Cocinetas Basin
upper Cenozoic sedimentary rocks. The oldest preserved The Cocinetas basin is a low area enclosed by the
structures in the Guajira Peninsula are a series of Serranias de Cocinas, Jarara, and Macuire. The basin
southeast-verging thrusts that place allochthonous over opens to the southeast where it extends with southeast-
autochthonous Cretaceous rocks. ward dip into the Golfo de Venezuela (Renz, 1956; Bürgl,
Falcón Basin—The Falcón basin is an east-west 1960; Rollins, 1965; Thomas, 1972a,b) (Figure 3). Tertiary
depression developed in northwestern Venezuela. sedimentary rocks rest unconformably on Cretaceous
Subsidence related to both extensional opening between metamorphic and sedimentary rocks. The total thickness
wrench fault systems and flexural loading by an of outcropping strata is more than 2500 m (Rollins, 1965),
advancing thrust front resulted in the deposition of but the Guajira-1 well penetrated less than 1000 m of
more than 6000 m of Oligocene and younger clastics and Cenozoic strata (Figures 3, 4).
reefal limestones. The area surrounding the Falcón basin
displays three structural styles. The first is found in the Cycle D (upper Eocene)—Sedimentation in the basin
stable highs of the Dabajuro platform and Paraguaná started in the late Eocene with the Macarao Formation.
Peninsula, where there is only minor deformation of a This unit is exposed only in small topographic lows along
thin Cenozoic cover over a Mesozoic metamorphic the eastern margin of the Serrania de Cocinas. In the type
basement. The second style is represented by regional section, the Macarao Formation is 240 m thick and
northwest-southeast striking growth faults developed contains gypsiferous silty shale, fine-grained glauconitic
between the major highs (Golfo de Venezuela, Urumaco sandstone with shale interbeds, and massive limestone
trough, offshore Aruba, and eastern La Vela Bay) (Thomas, 1972a,b). The age of this formation is still
(Payne, 1951; Audemard and Demena Arenas, 1985; debated, but the bivalve fauna includes the diagnostic
Boesi and Goddard, 1991). Changes in stratigraphic species Venericardia (Venericor) guajirensis and Turritella
thickness along these faults indicate that they were chira, indicating a late Eocene age (Stainforth, 1962a;
active during Oligocene–Miocene sedimentation. The MacDonald, 1964; Rollins, 1965; Thomas, 1972a,b).
third style occurs along the Falcón anticlinorium which
forms the inverted part of the basin. The anticlinorium is Cycle C (upper Oligocene–lower Miocene)—The
formed by elongated east-west folds superimposed on Siamana (or Sillimana) Formation lies unconformably on
previous tectonic styles. The contact between the anticli- top of the Macarao Formation. It is exposed around the
norium and the flat-lying strata of the coastal plain is margins of the basin, especially in the northwest between
marked by a zone of north-verging thrusts and asym- the Serranias de Cocinas and Jarara. The Siamana
metric and overturned folds, which are well exposed Formation consists of a lower conglomeratic member
south of Coro. and an upper limestone member (Renz, 1960). The upper
Prominent features of the Falcón basin are the east- limestone prograded across the underlying unit, and in
west right-lateral wrench faults. The most conspicuous the western part of the basin, it directly onlaps Jurassic
760
Macellari

Figure 2—Upper Eocene–Recent stratigraphic chart for the Guajira Peninsula, Falcón basin, and Paraguaná areas.
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 761

and Cretaceous rocks (Renz, 1960). The upper interval


consists of reefoidal limestone with bioclastic debris.
These biostromes are associated with the margins of the
relatively uplifted highlands, where they apparently
formed fringing reefs. In the subsurface, this twofold
division is less obvious, and most of the Siamana
Formation is composed of limestone with sandstone
interbeds. The thickness of this transgressive formation
varies from 100 m in the Guajira-1 well to 325 m at Cerro
Jimol (locality 1, Figure 3) (Renz, 1960). The upper
Siamana Formation contains microfauna diagnostic of
the Globigerina ciperoensis zone (Rollins, 1965), although
Duque-Caro (1974) considers the unit to be slightly
younger.
Overlying the Siamana is the Uitpa Formation. The
contact between the two is discordant, especially around
the edges of the basin (Rollins, 1965; Thomas, 1972b), but
it tends to be transitional in the center (Rollins, 1965). The
Uitpa Formation is composed of silty, selenitic clays and
shales, with an abundant microfauna. Fine grained,
calcareous sandstone interbeds are common in the lower
Figure 3—Cenozoic basins of the Guajira Peninsula,
showing location of wells, total Cenozoic isopachs in and upper parts of this formation (Thomas, 1972b). The
meters, and location of cross sections in Figures 5 and 6. thickness of the Uitpa ranges from 340 m near Uitpa to
Numbers in the Cocinetas basin show location of sites 200 m north of Siamana (Figure 3, localities 2 and 4,
referenced in Figure 4 and in the text: 1, Cerro Jimol; 2, respectively) (Rollins, 1965). The Uitpa Formation repre-
Quebrada Uitpa; 3, Guararies; 4, Siamana-Samuludo; 5, sents the maximum encroachment of the Cenozoic sea in
Guajira-1 well; 6, Bahia de Tortugas. the Guajira Peninsula. This formation is believed to have

Figure 4—Stratigraphic correlation of measured sections and well in the Cocinetas basin. See Figure 3 for locations.
762 Macellari

Figure 5—Correlation of wells in offshore Chichibacoa basin. Only the resistivity logs are shown. See Figure 3 for location.

an early Miocene age and includes the Globorotalia kugleri, has strong affinities with the Cantaure Formation of the
Catapsydrax dissimilis, Catapsydrax stainforthi, and lower Paraguaná Peninsula, indicating a middle–late Miocene
Globigerinatella insueta zones (Thomas, 1972b; Duque- age (Thomas, 1972a,b).
Caro, 1974).
Chichibacoa and Portete Basins
Cycle B (upper Miocene)—The contact of the Uitpa These basins extend north of the Serranias de Macuire
and Jimol formations has been described as gradational and Jarara (Chichibacoa) and north of the Serrania de
(Renz, 1960; Rollins, 1965; Thomas, 1972b). However, Cocinas (Portete) (Figure 3). Both basins merge and
Hunter (1974) recognized an abrupt lithologic break continue into the southern Caribbean Ocean. They
between the two. Duque-Caro (in Thomas, 1972a) also display a progressive northward increase in sedimentary
noted a marked hiatus between the top of the lower thickness to more than 4000 m.
Miocene and the base of the upper Miocene on the basis
of foraminiferal zones. A marked downward shift of Cycle D (upper Eocene)—Onshore, this cycle is repre-
coastal onlap is observed at the base of the Jimol sented by the Nazareth Formation (Renz, 1960; Stainforth,
Formation on geologic maps. Lithologically, the Jimol 1962b). It is exposed in a discontinuous belt north of the
Formation is composed of a calcareous sandstone or Serrania de Macuire. Two lithologic units are distin-
sandy limestone with some clay interbeds (Rollins, 1965; guished in the type section: a 30-m-thick lower unit
Thomas, 1972b). The shallow marine molluscs in these consisting of conglomerate, calcareous sandstone, and
beds indicate a middle Miocene age (Rollins, 1965; limestone, and an upper unit comprising 35 m of pebbly
Thomas, 1972b). Duque-Caro (in Thomas, 1972a) placed limestone with abundant marine macrofossils and
the Jimol Formation in the early late Miocene Globorotalia interbedded sandstones (Thomas, 1972a,b). The maxi-
menardii zone. This unit is over 600 m thick east of Uitpa mum thickness of the Nazareth Formation is 290 m. The
(locality 2, Figure 3), but thins rapidly toward the old contact with the Siamana Formation is believed to be
shoreline to the west (Rollins, 1965). unconformable because the two formations have different
The uppermost Tertiary of the Cocinetas basin is dips. The age of the Nazareth is interpreted as late Eocene
represented by the Castilletes Formation. This formation on the basis of foraminifera (Nummulites jacksoniensis) and
was deposited conformably on the Jimol, but it onlaps the echinoid genus Oligopygus (Thomas, 1972b). Contem-
older Cenozoic strata along the margins of the basin. poraneous strata are present in the Uashir, Puerto
Rollins (1965) distinguished two lithologic units in the Estrella, and Chimare wells (Figures 3, 5, 6). In the
Castilletes Formation: a lower limestone and an upper Chimare well, cycle D rocks overlie a metamorphic
claystone, with a combined thickness of 690 m. However, basement and include 48 m of shale. These sedimentary
the formation thins to less than 100 m to the west rocks become coarser and thicker toward the south
(Thomas, 1972b). The rich macrofauna of the Castilletes (Puerto Estrella well).
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 763

Figure 6—Interpreted seismic line in offshore Chichibacoa basin. See Figure 3 for location.

Figure 7—Interpreted regional seismic line through northwestern Falcón basin. Note major fault generating the Urumaco
trough.

Cycle C (upper Oligocene–lower Miocene)—This cycle about 1050 m. There are no age-diagnostic fossils in this
corresponds to the outcropping Siamana and Uitpa unit (Bürgl, 1960). In the subsurface, cycle B is wide-
formations. The thickness of cycle C ranges from less spread and is composed of shale and mudstone with
than 200 m in the south to over 1000 m in the north. The minor intercalations of limestone and rare coal (Santa
Uitpa Formation is probably slightly older in the Portete Ana well) (Figure 5). The subsurface thickness of cycle B
basin than in the adjacent Cocinetas basin, extending ranges from a few hundred meters at the southern end of
from the Globigerina ciperoensis ciperoensis zone (latest the basin to a maximum of 2290 m in the Punta Gallinas
Oligocene) to the Globigerinatella insueta zone (late early well. Its age is believed to range from middle to late
Miocene) (Thomas, 1972b). Miocene.
Within the Chichibacoa basin, available seismic data
indicate that cycle C is composed of three packages (C3, Cycle A (Pliocene)—Cycle A sedimentary rocks are
C2, and C1). At the base, C3 was deposited throughout exposed only along the northern edge of the Guajira
the basin, except for the eastern Chimare high, where the Peninsula, where they include solution-pitted sandy
Santa Ana well shows that Cenozoic sedimentation limestones that unconformably overlie the Taroa
began slightly later (Figure 5). This unit is composed of Formation (Thomas, 1972b). This 30-m-thick sequence is
sandstone with local coal seams and subordinate shales. included in the Gallinas Formation of Thomas (1972b). In
Seaward, these strata grade into shale (Chimare-1) and the subsurface, cycle A is more argillaceous and contains
limestone (Punta Gallinas-1). Subcycle C2 was deposited mudstone with subordinate sandstone and limestone. A
only in the deepest part of the basin. These sedimentary total of 580 m of Pliocene strata were penetrated in the
rocks, dated in the wells as late Oligocene, were appar- San Jose-1 well (Figure 5).
ently deposited after the pronounced late Oligocene sea
level drop (within the Globorotalia opima opima zone; Vail Falcón Basin
and Hardenbol, 1980). This interval is composed of
limestone in the Punta Gallinas-1 well and sandstone in The Falcón basin is an ENE-WSW elongated uplift,
the more proximal facies of the Uashir-1 well. with a core of older Oligocene strata and alkaline basaltic
The younger subcycle C1 was deposited throughout intrusions of Oligocene–Miocene age (K-Ar age of 22.9 +
the basin with fairly uniform thicknesses (770–1150 m). It 0.9 m.y.; Muessig, 1978, 1984) that are succeeded to the
consists of shale and interbedded sandstone and north by progressively younger Cenozoic strata. The
limestone attributed to the maximum transgression. basin contains a thick Oligocene–Miocene sequence that
Subcycle C1 equates with the lower Miocene Uitpa shale. was deposited in a series of NNW-SSE and NNE-SSW
trending troughs (Figure 7) (Wheeler, 1963; Muessig,
Cycle B (middle–upper Miocene)—In outcrop, cycle B 1984; Boesi and Goddard, 1991). Rocks of the Falcón anti-
is included in the Taroa Formation, which rests discon- clinorium are strongly folded and overturned along
formably on the Uitpa Formation. It is composed princi- north-verging thrusts. This deformation is attributed to
pally of sandy limestone, calcareous sandstone, pebbly transpressive displacement of the Oca fault (Figure 1).
conglomerate, and subordinate mudstone (Thomas,
1972b). These rocks cap the mesas along the northern Cycle D (upper Eocene)
and northeastern coasts of the Guajira Peninsula. The Sedimentation in the Falcón basin probably started in
maximum measured thickness of the Taroa Formation is the late middle Eocene in western Falcón and as late as
764 Macellari

Figure 8—Oligocene paleogeography of the Falcón basin (cycle C). (Modified from Wheeler, 1960.)

early late Eocene in eastern Falcón. Sediments related to Cycle C (Oligocene–lower Miocene)
the initial formation of the basin in western Falcón are Sedimentation in the Falcón trough began during early
represented by the Agua Negra Group of upper middle late Oligocene time (Wheeler, 1963; Diaz de Gamero,
Eocene age (Figure 2). These rocks are included in the 1977a, 1985). These rocks form a suite of facies deposited
Santa Rita, Jarillal, and La Victoria formations (Guevara, during a marine transgression that was punctuated by a
1967). The Santa Rita Formation contains 400–500 m of regressive episode near the Oligocene–Miocene
coarse conglomerate, sandstone, and marl with boundary. The regressive episode separating subcycles
interbedded limestone (Guevara, 1967). Its lower contact C2 and C1 is represented by a widespread sheet of coarse
is not well exposed, but it is usually in fault contact with clastics. Cycle C rocks are well exposed in the Falcón anti-
older Eocene strata (Gonzalez de Juana et al., 1980). clinorium. They were deposited on Eocene strata to the
These strata are conformably overlain by the Jarillal west, on metamorphic Cretaceous rocks to the east, and
Formation, up to 375 m of dark shale with sandstone on Paleocene–Eocene turbidites in the center of the basin.
intercalations (Guevara, 1967). The invertebrate fauna Five different areas of deposition are recognized in the
and palynomorphs indicate a middle Eocene age Falcón basin (Figures 2, 8).
(Wheeler, 1960; Hunter, 1974, 1978).
The upper part of the Agua Negra Group is the La Central Area—The central area is the main depocenter
Victoria Formation, comprising 2400 m of dark shale with of the Falcón basin. Sedimentation started with the El
interbedded sandstone and coal (Sutton, 1946; Guevara, Paraiso Formation, the lower part of which is a 205-m-
1967). The La Victoria Formation was deposited predomi- thick chert-rich sandstone with shale interbeds (Wheeler,
nantly in a lacustrine environment with an incursion of 1963; Diaz de Gamero, 1977a), followed by a quartz and
brackish water (Sutton, 1946; Gonzalez de Juana et al., chert pebble conglomerate (Gonzalez de Juana et al.,
1980). 1980). The upper member of the El Paraiso Formation
In eastern Falcón, Eocene rocks are more than 400 m consists of 330–450 m of dark shale with interbedded
thick and are included in the dark calcareous shales of quartzose sandstone. This interval becomes more arena-
the Cerro Misión Formation (Figure 2) (Renz, 1948). The ceous toward the south. The upper El Paraiso is inter-
shales contain microfossils of the Globigerinatheka semiin- preted as a regressive sequence that prograded toward
voluta zone (late Eocene) (Hunter, 1974). The Cerro the north and northeast (Figure 8).
Mision disconformably overlies middle Eocene rocks An impoverished microfaunal assemblage dominated
(Hunter, 1972, 1974; Diaz de Gamero, 1985). by arenaceous foraminifera occurs in this unit. The taxa
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 765

Figure 9—Early Miocene paleogeography of the Falcón basin (subcycle C1). (Modified from Wheeler, 1960; Gonzalez de
Juana et al., 1980.)

suggests brackish depositional environments. The upper Miocene; Diaz de Gamero, 1977b).
portion of this formation has fauna diagnostic of the During the early Miocene, subcycle C1 was deposited
Globorotalia opima opima zone (middle late Oligocene); the (Figures 2, 9). Regional subsidence resulted in onlap and
lower interval may be included in the Globigerina amplia- overstepping of the previous margins of the basin. In the
pertura zone (early late Oligocene) according to Diaz de center of the basin, sedimentation started with shales of
Gamero (1977a,b). the Agua Clara Formation, a sequence of dark ferrugi-
A thick sequence of dark silty shale in the center of the nous shale with intercalations of silty and calcareous
Falcón basin is included in the Pecaya Formation (Diaz sandstones that are locally glauconitic (Gonzalez de
de Gamero, 1977a,b). Sandstone and limestone intervals Juana et al., 1980). The shales are usually very fossilif-
are more numerous toward the margins of the basin erous and locally have a petroliferous odor. The lower
(Wheeler, 1963). Within the Pecaya, the San Juan de la part of the Agua Clara Formation contains the Cauder-
Vega Member consists of shales with interbedded alito Limestone, a richly fossiliferous unit containing
sandstone, siltstone, and minor limestone (Figure 2). The corals, bryozoans, bivalves, gastropods, and large
sandstones contain shell fragments and are strongly foraminifera. The thickness of the Agua Clara varies
bioturbated (e.g., Ophiomorpha). The San Juan de La Vega from 1320 to 1600 m (Wheeler, 1963). Because this unit is
Member tapers to the west and east and was probably preserved only along the margins of the basin, the
sourced from the Dabajuro high (Figure 8). The biota molluscs and foraminifera indicate shallow water envi-
indicate shallow marine deposition within the deeper ronments (Diaz de Gamero, 1977a). The age of this unit
water Pecaya shale (Diaz de Gamero, 1977a). This unit spans the early Miocene Catapsydrax dissimilis to Praeor-
reflects lowstand deposition possibly associated with the bulina glomerosa zones (Diaz de Gamero, 1977b).
pronounced worldwide drop in sea level that took place Along the southern and southwestern margins of the
in the Globorotalia opima opima zone (see Vail and Sierra de San Luis, subcycle C1 was initiated with the
Hardenbol, 1980). Pedregoso Formation (Figures 2, 9). This shale,
The Pecaya Formation is 600 m thick in central Falcón, interbedded sandstone, and bioclastic limestone have
but it thickens westward to 1300 m. The lower part of the been interpreted as proximal turbidite deposits (Diaz de
Pecaya contains a fauna diagnostic of the Globorotalia Gamero, 1977a). This unit ranges in thickness from 120 m
opima opima zone (middle late Oligocene), whereas the in the east to 940 m in the west (Wheeler, 1963). The rich
upper part includes taxa associated with the Globigeri- microfauna indicates the early Miocene Globigerinoides
noides primordius and Catapsydrax dissimilis zones (early primordius and lower Catapsydrax dissimils zones
766 Macellari

(Gonzalez de Juana et al., 1980). This formation was is called the Agua Salada (Diaz de Gamero, 1985) or
probably deposited in a deep water environment that Hueque depocenter (Wheeler, 1963). South of this high,
shallowed toward the top (Diaz de Gamero, 1977a). strata are considerably coarser grained and are included
in the Guacharaca and Casupal formations (Renz, 1948;
Western Falcón—In the western Falcón basin, the Blow, 1959; Diaz de Gamero, 1985) (Figures 8, 9).
Pedregoso Formation is more arenaceous and thicker
and correlates with the Castillo Formation (Figures 2, 8) Cycle B (middle–upper Miocene)
(Wheeler, 1960, 1963). The Castillo is conformable on the
At the end of the early Miocene, the pattern of sedi-
basinal facies tract of the Pecaya shale. Farther west, it
mentation in the Falcón basin changed significantly as a
has an angular contact with Eocene strata (Figure 2). The
result of a relative drop in sea level or rapid uplift associ-
Castillo is dominated by sandstone and conglomerate,
ated with the early stages of deformation. After this
with lesser amounts of shale and coal seams (Wheeler,
event, renewed subsidence initiated a new cycle of sedi-
1963). The unit was deposited under marine conditions
mentation, which at first blanketed the entire basin,
toward the center of the basin, with brackish and even
including topographic relief such as the Dabajuro
continental facies influencing the western margin. The
platform and the Paraguaná Peninsula, which lacked a
main depocenter of the Castillo Formation (up to 1480 m)
cycle C cover. Newly uplifted areas to the south
forms a belt parallel to the old western coastline. Lacking
provided considerable sedimentary input. Cycle B sedi-
index fossils, the age of the Castillo is estimated as late
mentation occurred in shallow or restricted environ-
Oligocene–early Miocene on the basis of its stratigraphic
ments, except in the northeastern part of the basin (Agua
relationships (Gonzalez de Juana et al., 1980).
Salada) and the northern Paraguaná Peninsula (V. F.
Coro High—The reefal complex of the Sierra de San Hunter, 1986, personal communication) where deep
Luis developed along the southern margin of the Coro water facies persisted.
high during Oligocene–Miocene time (Figures 8, 9).
While the core of the reef is included in the San Luis Central Area—In the center of the Falcón basin,
Formation, forereef and open marine facies of the Pedre- cycle B starts with the Cerro Pelado Formation (Figures
goso Formation were deposited southward. In contrast, 2, 7, 10). This formation consists mainly of laminated
peri-reefal facies interfingered with deltaic and coastal sandy shale and interbedded fine-grained sandstone
shale and sandstone facies of the Patiecitos Formation, (Liddle, 1946). The sandstones are usually cross bedded
and with sandstone and conglomerate facies of the and contain ripple cross lamination and lignite seams up
Guarabal Formation along the northern margin to 1 m thick (Gonzalez de Juana et al., 1980). The Cerro
(Gonzalez de Juana et al., 1980). Pelado was deposited unconformably on top of the Agua
The San Luis Formation consists of massive reef lime- Clara shale. It is about 1000 m thick and is replaced to the
stones containing large foraminifera, algae, and locally east by shales of the Agua Salada Group. Diaz de
large brain corals and bivalves (Wheeler, 1963). The Gamero (1977b) assigned the Cerro Pelado to the Globoro-
maximum growth of the reef is restricted to the eastern talia foshi peripheroronda zone (early middle Miocene).
part of the Serrania de San Luis, where reefal limestones The overlying Socorro Formation crops out in the hills
are more than 100 m thick (Diaz de Gamero, 1977a). The of western Falcón and in the Cumarebo zone. In the type
minimum thickness of the San Luis Formation is section, near the town of Socorro, the Socorro has a basal
estimated to be 800 m (Wheeler, 1960), and its age is shale (Gonzalez de Juana et al., 1980) and an upper
inferred to be late Oligocene (Globorotalia opima opima sandstone interval. The upper fine-grained sandstones
zone) to early Miocene (Catapsydrax stainforthi zone; are oil reservoirs in the Cumarebo field. The Socorro
Gonzalez de Juana et al., 1980). Formation is 1570 m thick but thins to 660 m south of
Coro (Gonzalez de Juana et al., 1980). The microfauna of
South-Central Falcón—The south-central Falcón the lower part of the succession is assigned to the
basin also developed carbonate facies along the southern Globorotalia foshi foshi zone (earliest middle Miocene)
extension of the Coro high (Figures 8, 9). The entire (Diaz de Gamero, 1977b). The Socorro Formation was
Oligocene–lower Miocene sequence (cycle C) consists of deposited in a coastal environment to the west, which
a series of ridge-forming limestone, sandstone, and shale became gradually more open marine to the east
up to 125 m thick, which together form the Churuguará (Gonzalez de Juana et al., 1980). The Socorro succession
Formation (Wheeler, 1963). The Churuguará has a grades upward into the Urumaco Formation in the
conformable contact with the Jarillal Formation below Urumaco trough area and into the Caujarao Formation
and the Agua Clara Formation above. Foraminifera in farther east.
the lower part of the section are indicative of the Globoro- The Caujarao Formation includes 1220 m of shale and
talia opima opima zone (middle late Oligocene), whereas interbedded marl, limestone, and fine-grained sandstone
the upper part contains early Miocene foraminifera that are oil bearing in the Cumarebo field (Payne, 1951).
(Wheeler, 1963). Rapid facies changes are reflected in a diverse strati-
graphic nomenclature (see Gonzalez de Juana, 1937;
Eastern Falcón—In the eastern area of the Falcón Petzall, 1959; Payne, 1951; Senn, 1935; Vallenilla, 1961;
basin, deposition during cycle C also formed a variety of Ministerio de Minas, 1970; Gonzalez de Juana et al.,
facies. North of the Esperanza-Guacharaca high, deep 1980). The Caujarao Formation was deposited on a
water shales (Agua Salada Group) accumulated in what marine platform in a warm, tropical sea (Petzall, 1959).
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 767

Figure 10—Middle–late Miocene paleogeography of the Falcón basin (cycle B). (Modified from Wheeler, 1960; Gonzalez de
Juana et al., 1980.)

The rich fauna indicates an age ranging from the Globoro- have occurred in a piedmont-coastal area, with a pro-
talia menardii zone (late middle Miocene) to the Globoro- venance located to the south (Gonzalez de Juana et al.,
talia acostaensis zone (late Miocene) (Diaz de Gamero, 1980). On the Dabajuro Platform, La Puerta strata rest
1977b). directly on Eocene rocks. The maximum thickness of this
Toward the west, in the area of the Urumaco trough, unit (2140 m) was encountered in the QMC-1X well
the Caujarao is replaced by the Urumaco Formation (Figure 11).
(Figure 10). These are more proximal facies than the
Caujarao Formation and include marine and nonmarine Northeastern Falcón—In the northeastern Falcón
shale, sandstone, and coal seams. The thickness of the basin, sediments deposited during cycle B are included
Urumaco Formation is about 1675 m (Ministerio de in the upper part of the Agua Salada Group (north) and
Minas, 1970; Gonzalez de Juana et al., 1980). the Pozón Formation (south) (Figure 10). In the Agua
La Vela Formation is conformable on the Caujarao Salada depocenter, the base of cycle B is represented by
Formation in the Coro area. It includes 590 m of the El Salto Formation. According to Diaz de Gamero
calcareous sandstone with shale interbeds, reflecting (1985), this unit has four lithofacies: (1) olistostrome
beach and estuarine depositional environments blocks 1–5 m in diameter of sandstone, siltstone, and
(Gonzalez de Juana et al., 1980). The age of the La Vela coal, encased in shale; (2) fine-grained sandstone in beds
Formation ranges from the Neogloboquadrina dutertrei 20–50 cm thick; (3) massive siltstone with plant remains
zone to the Globorotalia margaritae zone (late at the base coarsening upward into sandstone; and (4)
Miocene–early Pliocene) (Diaz de Gamero, 1968). The La thick sandstone beds. The El Salto is estimated to be
Vela Formation is succeeded to the east by more open 800–900 m thick. According to Diaz de Gamero (1985),
marine facies of the El Veral Formation and to the west the El Salto Formation is a deltaic deposit related to
by more continental facies of the Codore Formation. growth faulting and turbidity current processes. These
deltaic sediments encapsulated in an otherwise deep
Western Falcón—Sediments deposited in the western water Agua Salada sequence are attributed to a sea level
Falcón basin during cycle B are included in the La Puerta lowstand. The age of this unit ranges from the Praeorbu-
Group or Formation (Figures 2, 7, 10, 11). This unit lina glomerosa to the Globorotalia foshi peripheroronda zone
comprises claystone, sandstone, and lignite overlain by (Diaz de Gamero, 1985).
massive and cross-bedded sandstone that is capped by South of the Agua Salada depocenter, the upper part
varicolored shales (Halse, 1937). Deposition is believed to of the sequence contains 150–300 m of shallow water
768 Macellari

Figure 11—Correlation of wells in western Falcón basin across the Oca fault.

carbonates of the Capadare Formation (Figure 10). of the uplifted Falcón anticlinorium (San Gregorio and
Capadare sedimentation started in the early middle Coro formations). In the northeastern coastal area, sedi-
Miocene (Globorotalia foshi foshi zone) and continued into mentary rocks are composed of calcareous clay, fossilif-
the late middle Miocene (Diaz de Gamero, 1985). erous marl, and argillaceous limestone (Punta Gavilán
Formation).
South-Central Falcón—In south-central Falcón basin,
cycle B rocks are represented by the Pozón Formation.
This is a shaly succession with glauconitic sandstones Paraguaná Peninsula
and marls at the base (Renz, 1948). Reworked The Paraguaná Peninsula extends north of the Falcón
foraminifera at the base suggest that the Pozón formation basin into the Caribbean Sea. Structurally, the peninsula
was locally preceded by erosion (see Renz, 1948). The is the northern part of a NNW-SSE trending high.
Pozón Formation is up to 1040 m thick. Micropaleon- Cenozoic sedimentation started at least by the late early
tology assigns these rocks to the Globigerinatella Miocene (cycle B), but seismic data suggest that early
insueta–Globigerina bulloides zones (early middle Miocene and even late Oligocene sediments might have
Miocene–late Miocene; Diaz de Gamero, 1985). accumulated in more restricted areas north of the
peninsula (Figures 2, 8, 9, 10, 12).
Southeastern Casupal—In the southeastern Casupal The Cenozoic sequence rests on metamorphic and
depocenter, Casupal sandstone and conglomerate are igneous basement. These sediments surround igneous
succeeded by 1300 m of shale, clay, and limestone of the and metamorphic highs in a series of concentric cuestas
Agua Linda Formation (Ministerio de Minas, 1970) dipping seaward up to 10°. Regional seismic lines show
(Figures 2, 10). The Agua Linda is conformable on the that the thickness of the Cenozoic stratigraphy increases
Casupal Formation and is overlain disconformably by progressively away from the center of the Peninsula
the Capadare limestone (Gonzalez de Juana et al., 1980). where the basement is exposed. In the southern part of
the peninsula, the Manuel-1 and Cardón-1 wells encoun-
Cycle A (Pliocene) tered 1067 m and 1159 m of Cenozoic strata, respectively
A major phase of basin inversion took place during (Figures 8, 9, 10). Manuel-1 bottomed in red beds of
the Pliocene. Coarse clastic facies were deposited north unknown age (possibly late Oligocene or early Miocene)
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 769

Figure 12—Interpreted northwest-southeast seismic line, Paraguaná Peninsula. The “Paraguaná reflector” is a high-
amplitude event at, or close to, the base of the sedimentary sequence. This reflector probably represents a volcanic body,
most likely a sill, since it appears to be slightly discordant with other stratal surfaces.

and basalt. The sedimentary thickness is much greater in


the northern peninsula (up to 4000 m), where antithetic
growth faults, active during much of the Miocene,
formed significant depocenters (Figure 12).
The outcrops of cycle B on the Paraguaná Peninsula
start with the Cantaure Formation. These are fossiliferous
silty shales with thin layers of gypsum and minor
sandstone and limestone intercalations (Hunter and
Bartok, 1974). The Cantaure Formation is up to 75 m
thick. An early Miocene age is inferred (Gobigerinatella
insueta to Praeorbulina glomerosa zones; Hunter and Bartok,
1974), possibly extending into the middle and late
Miocene (V. F. Hunter, 1986, personal communication).
The overlying Paraguaná Formation comprises most
of the sedimentary rocks exposed on the peninsula. It
consists of fossiliferous shale, siltstone, and limestone
(Rodriguez, 1968; Hunter and Bartok, 1974). The fauna
indicates the Globorotalia margaritae zone (Pliocene)
(Hunter and Bartok, 1974).

La Vela Bay
La Vela Bay is an offshore extension of the Falcón
basin and contains a similar stratigraphy. It displays a
regional tilt to the south and southeast, where more than
4500 m of strata have accumulated. The eastward
increase in thickness is controlled by eastward-dipping
growth faults (Figure 13). Seismic and well information
indicate the presence of cycles C, B, and A (lower
Miocene–Recent). The contact between cycles C and B
forms a persistent reflector (reflector “C”) (Biju-Duval et
al., 1979) or a regional unconformity (Gonzalez de Juana
et al., 1980). Figure 13—Structure at the top of the Cauderalito
Thirteen exploration wells and a dense seismic grid limestone and location of wells in the La Vela Bay (from
make La Vela Bay, one of the best known of the southern Corpoven).
Caribbean basins, although much of this information is
still confidential. The basement, reached in eight wells,
consists of gneiss, phyllite, and metamorphosed igneous
rocks, dated radiometrically as Cretaceous (Gonzalez de
Juana et al., 1980). In the southwestern part of the bay,
770 Macellari

Figure 14—Interpreted seismic line in offshore Aruba. Note the conspicuous tilting of cycle D strata prior to deposition of
cycle C.

Cenozoic sedimentation began with a continental red the Casupal (south) and Agua Salada (north) depocen-
bed sequence of shale, siltstone, and fine-grained to ters. The former is characterized by a predominantly
conglomeratic arkosic sandstone (Gonzalez de Juana et coarse clastic fill, whereas the latter contains more argilla-
al., 1980). An Oligocene age is inferred for these beds ceous facies.
(Vasquez, 1975). The Cayo Sal-1 well (Figure 8) penetrated a middle
Overlying this red bed sequence are the sandstones Eocene–Quaternary section. A basal middle–upper
and dark claystones of the Guarabal Formation. These Eocene shale is followed by a 1900-m-thick cycle C unit
strata form discontinuous pockets restricted to relatively (Oligocene–lower Miocene), composed of sandstone,
small depocenters that have a maximum thickness of shale, and limestone in an overall upward-fining succes-
550 m (Corpoven, 1980, proprietary report). The sion. Cycle B (lower Miocene–lower Pliocene) includes
Guarabal is the backreef facies of the San Luis reefal 1450 m of claystone and sandstone capped with
limestone tract. This formation has been interpreted as an limestone. The youngest strata encountered are 380 m of
important source of oil in the La Vela Bay area clay and limestone of Pleistocene age.
(Corpoven, 1980, proprietary report). The Guarabal is
followed by the Agua Clara Formation, consisting of the Aruba
basal Cauderalito limestone and calcarenite, followed by
dark shale with local thin limestone and calcarenite The island of Aruba is located on the westernmost
interbeds (Vasquez, 1975; Gonzalez de Juana et al., 1980). extension of the basement high that extends from La
These argillaceous rocks are also considered to be oil- Orchila to Aruba (Figure 1). Aruba has a core of Creta-
prone source rocks in the area (Boesi and Goddard, ceous igneous and metamorphic rocks, surrounded by a
1991). rim of thin upper Cenozoic deposits. Seismic surveys
The base of cycle B is formed by the Socorro west and southwest of Aruba show a thick sedimentary
Formation, an 850–1740-m-thick claystone succession sequence (Figure 14). Three seismic lines were available
with intercalations of calcarenite and fine- to medium- for this study. Elsewhere on the southwestern Caribbean
grained sandstone. This is followed by 260–580 m of margin, the reflector that separates cycles B and C
glauconitic calcarenite with interbedded medium- (reflector “C” of Biju-Duval et al., 1979) is interpreted as a
grained to conglomeratic sandstone of the Caujarao late early Miocene unconformity (Hunter, 1974, 1986).
Formation. The upper part of the cycle B section is the La The ages of younger unconformities are based on corre-
Vela Formation, comprising mudstone, calcarenite, and lation with the global sea level chart of Vail and
fine-grained sandstone. The uppermost rocks of La Vela Hardenbol (1980).
Bay (cycle A) include fine-grained sandstone with subor- A rapid increase in sedimentary thickness is observed
dinate siltstone, claystone, and limestone. These beds are west of Aruba. This change is controlled by a northwest-
assigned a Pliocene–Pleistocene age and are 120–145 m southeast trending normal fault parallel to the south-
thick (Gonzalez de Juana et al., 1980). western Aruba coast. A maximum of 5000 m of Cenozoic
sedimentary rocks is observed, but common sedimentary
Offshore Eastern Falcón (Golfo Triste) thicknesses vary between 2500 and 3500 m. Based on
regional correlations, this stratigraphy is believed to
Cycles A, B, and C are observed in regional seismic overlie Cretaceous basement similar to that exposed in
lines of the Golfo Triste. The stratigraphy and structure Aruba.
are apparently a continuation of the trends along the The succession attributed to cycle D is up to 1400 m
eastern coast of Falcón (Figures 8, 9, 10). The extension of thick, but generally ranges from 150 to 450 m. These
the Cerro Mision–Chichiriviche high to the east separates rocks were block faulted and tilted prior to deposition of
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 771

the overlying cycle C (Figure 14). The large acoustic lines, well logs, and measured outcrop sections. Data for
impedance of the basal part of cycle D may indicate the the Bonaire basin were derived from seismic lines inter-
presence of well-lithified limestone similar to the preted by Biju-Duval et al. (1979).
Nazareth and Macaro formations of the Guajira Maximum sedimentary thicknesses are preserved
Peninsula. The Nazareth was also block faulted and both east and west of the Paraguaná Peninsula. Two
tilted before deposition of cycle C strata. These rocks are depocenters are observed in the Golfo de Venezuela area.
possibly equivalents of the Butuco Limestone, which is The western depocenter has up to 7500 m of Eocene
poorly exposed in Aruba. In the northwestern part of the strata, whereas the eastern depocenter has as much as
basin, this block-faulted sequence is associated with 7500 m of upper Oligocene–Miocene rocks. The
reflectors that are alternatively strong and weak and also depocenter located east of the Paraguaná Peninsula has
discontinuous laterally. These may be interpreted as a up to 7000 m of upper Oligocene–Pliocene strata. Discus-
continental sequence developed in the initial rifting of sion of the isopachs of cycles D, C, and B, as well as the
the basin. postcollisional geologic history of this area, are presented
Strata interpreted as equivalent to cycle C are widely here.
distributed offshore from Aruba, where they have a rela-
tively uniform thickness, indicating a period of regional Cycle D
subsidence. Maximum thickness of these rocks is about
2400 m. Cycle C rocks are correlated with the Siamana Cycle D (upper Eocene) overlies Mesozoic and
and Uitpa formations of the Guajira Peninsula and with Paleogene rocks related to the older Caribbean compres-
the El Paraiso–Pecaya–Agua Clara sequence of the sional event. This cycle includes the Macarao and
central Falcón basin. Nazareth formations in the Guajira Peninsula, formed by
Deposits attributed to cycle B are distributed erosional debris derived from the nearby landmasses,
throughout the basin and increase in thickness west and followed by deeper marine sedimentation. These
southeast of Aruba. The thickest section is preserved in deposits are confined to the northeastern and south-
the southeastern part of the basin, where up to 1800 m eastern borders of the peninsula (Figure 15). In the north-
are estimated. The succession offshore from Aruba is western Falcón basin, sedimentation was initiated with
believed to reflect a deep water section of shale and clastics of the Santa Rita and Jarillal formations. The
mudstone. At least four subcycles are recognized in the Paraguaná Peninsula was a positive feature during this
cycle B sequence, forming a series of prograding clino- time. In the eastern Falcón basin, cycle D was initiated
forms (Figure 14). These cycles are best defined near the with deposition of the Cerro Misión shale. Eocene sedi-
Aruba high and toward the southeastern part of the mentary rocks are also preserved in the Golfo Triste area.
Aruba basin. They reflect relative sea level fluctuations Up to 2000 m of continental and shallow marine strata
and substantial sediment influx. Similar progradational occur in the Aruba area. The nature of the rocks in the
wedges are observed in La Vela Bay. Cycle B is possibly Bonaire basin is unknown, but they are believed to
equivalent to the Socorro, Caujarao, and La Vela forma- consist of shale and turbidites.
tions of La Vela Bay and of the north-central part of the Subsidence during cycle D was controlled by progres-
Falcón basin. Pliocene–Recent cycle A sediments increase sive extension of an already discontinuous landmass.
in thickness away from Aruba, ranging from 100–300 m. Paleohighs outlined rhomboidal depocenters that were
Three dry wildcat exploration wells recently drilled in closer before the initiation of late Eocene extension of the
the Aruba region support these stratigraphic relation- southern Caribbean margin. This geometry suggests that
ships. Curet (1992) documented three sequences that are the Paleocene Serie Domi conglomerate of Bonaire was
separated by regional unconformities. The lower shed from the Paraguaná Peninsula. The complex iso-
sequence, apparently correlative of cycles D, C3, and C2, pachs in the western Falcón basin are the result of right-
ranges in age possibly from the late Eocene to the latest lateral displacement of previous NNW-SSE trending
Oligocene (Globorotalia kugleri zone). The sequence is depocenters by the Oca and associated faults (Feo-
composed of varicolored mudstone followed by clay- Codecido, 1972).
stone with a limestone cap. The middle sequence, equiv-
alent to cycle C1, is composed of deep water claystone Cycle C
and shale of early–middle Miocene age. The upper
sequence, equivalent to cycles B and A, consists of coars- Early Oligocene tectonic activity followed deposition
ening-upward sequences of soft claystone with of cycle D (Figure 16). In the northern Guajira Peninsula,
sandstone interbeds of late Miocene–Pleistocene age. deformation of the Macarao Formation and tilting of the
Nazareth Formation took place during this event. This
was probably also the time of tilting of cycle D rocks in
REGIONAL ISOPACHS AND GEOLOGIC the southern part of the Aruba basin. In the Falcón basin,
this event is marked by a regression and uplift of the
HISTORY southern margin of the basin.
Isopachs of postcollisional strata (Eocene and The late Oligocene was marked by high subsidence
younger) deposited in northwestern Venezuela, the rates and deep water sedimentation throughout the
Guajira Peninsula, and offshore Aruba are shown in study area. The pull-apart basin was well developed and
Figures 15, 16, and 17. They are based on regional seismic bounded to the north by the Aruba-Curaçao-Bonaire
772
Macellari

Figure 15—Isopach map of cycle D strata, upper Eocene.


Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 773

Figure 16—Isopach map of cycle C strata, upper Oligocene–lower Miocene.


774 Macellari

chain, to the south by the platform edge of the Falcón rupted offshore where up to 2000 m of argillaceous
basin, and to the west by the Dabajuro-Guajira platform. material was deposited.
Deposition occurred along two main trends. The first, The Falcón basin was progressively deeper toward
oriented NNW-SSE, was controlled by extension along the east. In the western part of the basin, the Dabajuro
preexisting cycle D troughs. The second, oriented ENE- platform was covered by deltaic and shallow marine
WSW, was aligned along the axis of the Falcón basin. deposits of the La Puerta Formation. These were progres-
Sedimentation in the Guajira Peninsula started with a sively replaced to the east by marine units, including the
basal conglomerate (Siamana Formation) eroded from Cerro Pelado, Socorro, and Caujarao formations, and
the uplifted Jarara, Macuira, and Cocinas highlands. finally by deep water black shales of the Agua Salada
These were followed by early Miocene shale deposited Group. During this time, the Paraguaná Peninsula was
during the maximum Cenozoic trangression, when most largely covered by shale and limestone of the Cantaure
of the Guajira Peninsula was drowned. Cycle C deposits Formation. These sediments were deposited in deep
in the Chichibacoa basin increase in thickness to more water north of the peninsula.
than 1000 m in an offshore direction. In the Cocinas
basin, these strata are usually thinner than 200 m. Cycle A
In the Falcón basin, the locus of sedimentation
migrated east of the cycle D depocenter. No cycle C The original distribution of cycle A has been inter-
deposits are preserved on the Dabajuro platform, which rupted by erosion, thus precluding the construction of an
had subsided rapidly during the previous cycle. Marked isopach. Three depocenters include the Urumaco trough
cycle C subsidence occurred along the axis of the Falcón in the west with up to 1400 m of strata, a depocenter
basin (ENE-WSW), where 3600–3800 m of section are north of the Paraguaná Peninsula with up to 1200 m of
preserved. Up to 3800 m of cycle C sediments were also deposits, and a depocenter in the Golfo Triste area with
deposited in the NNW-SSE trending Urumaco trough, up to 1800 m of Pliocene sedimentary rocks.
where subsidence was controlled by the westward- On land, this sequence is generally shallow marine to
dipping Sabaneta normal fault that forms the eastern fluviodeltaic. On the Guajira Peninsula, the Pliocene
margin of the trough. The southern margin of this Gallinas Formation was deposited after a period of
depocenter is presently offset about 15–20 km by the tectonism. In Falcón, this deformation resulted in basin
right-lateral Oca fault, suggesting that a large amount of inversion and uplift of the Falcón anticlinorium. In the
displacement along this fault occurred after the early lowlands of the Paraguaná Peninsula and in the coastal
Miocene. Cycle C rocks in the Falcón basin developed as area east of Cumarebo, cycle A rocks are composed of
a series of concentric belts ranging from coarse clastics calcareous clay and fossiliferous limestone of the
around the basin margins to deep water shales toward Paraguaná and Punta Gavilán formations, respectively.
the center. In the northern and southern parts of the The most active sedimentation during the Pliocene is
central Falcón basin, thick carbonate reefs developed on observed in the Golfo Triste area, where a thick sequence
regionally high areas. of clay and limestone was deposited.
Maximum transgression took place during the early
Miocene. During this time, the Agua Clara shale was
deposited throughout most of the Falcón basin, while
calcareous facies and reefs continued to develop in the SUBSIDENCE RATES THROUGH TIME
San Luis and Churuguará areas. The northern part of the
Paraguaná Peninsula and large areas of La Vela Bay To better understand the relationships between
were flooded for the first time. tectonics and sedimentation, 53 burial curves based on
well and seismic information were calculated for the
Guajira Peninsula, Falcón, and Aruba areas. These curves
Cycle B are not decompacted and do not include bathymetry.
By the middle–late Miocene, several of the blocks Representative curves are shown in Figure 18. Seven
present during cycle C were dissected and marine sedi- areas with common subsidence characteristics are recog-
mentation took place throughout the study area (Figure nized: northern Guajira Peninsula, Dabajuro platform,
17). The rapidly subsiding Urumaco trough and a Urumaco trough, Paraguaná Peninsula, eastern Falcón
subparallel depocenter east of the Paraguaná Peninsula and La Vela Bay, Aruba, and offshore eastern Falcón.
still maintained their own identities. Cycles C and B are
separated by a pronounced regional unconformity that Northern Guajira Peninsula
has a conspicuous seismic expression (Biju-Duval et al.,
1979, 1982; Gonzalez de Juana et al., 1980). Cycle B is Subsidence rates calculated from seven wells were
characterized in La Vela Bay and offshore Aruba by available for this area. Subsidence rates in the northern
sigmoidal clinoforms. Guajira were low during the late Eocene, with a mean
In the Cocinetas basin, cycle B contains offlapping value of 68 m/m.y., but with a large variation between
calcareous sandstone and sandy limestone of shallow localities (Figure 19). The Oligocene was a time of
marine origin (Jimol Formation). In the Portete and reduced subsidence, with mean rates varying between 49
Chichibacoa basins, there is a hiatus between cycles C and 54 m/m.y. A pattern of rapid subsidence was estab-
and B onshore, but sedimentation was almost uninter- lished during the early Miocene (cycle C1) and continued
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 775

Figure 17—Isopach map of cycle B strata, middle–upper Miocene.


776 Macellari

into the middle–late Miocene (cycle B). The mean rates


for the Miocene varied from 136 to 139 m/m.y. Subsi-
dence again decreased during the Pliocene (cycle A) with
consistent values close to 77 m/m.y. In summary,
maximum subsidence in the northern Guajira Peninsula
took place during the Miocene, with slower rates in the
Eocene, Oligocene, and Pliocene.

Northwestern Falcón Basin


Information on this area, broadly coinciding with the
Dabajuro platform, was obtained from six wells. Cycle C
(Oligocene–lower Miocene) sedimentary rocks are
scarce, so that cycle B rests directly on Eocene rocks.
(a) In some areas north of the Oca fault, Eocene deposits
are thin and wells (QMC-1X and QMD-1X) reached
basement (Figure 18). In general, this was an area of low
sedimentation rates, except for an Eocene rate of 170
m/m.y. (Figures 18, 19). However, this value is no more
accurate than the estimated age of the sequence. Only
two of the analyzed wells (Santa Cruz and AMN-2X)
record early Miocene C1 sedimentation (Figure 18).
Subsidence was moderate during the middle and late
Miocene, with a mean rate of 121 m/m.y., and was fairly
uniform throughout the area.
(b)
Urumaco Trough
Figure 18–(a) Selected subsidence curves for the study The Urumaco trough is characterized by high subsi-
area. All curves were derived from well data, with the dence rates during cycles C and B (Figures 18, 19). Four
exception of Aruba-6 and Paraguaná-12, which were calcu- subsidence curves were calculated. Correlation of seismic
lated from seismic lines. (b) Location map. information and well data indicates that rapid sedimen-

Figure 19—Average subsidence rates by area. The coefficient of variation indicates the degree of deviation of the data from
the calculated mean value.
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 777

tation probably started in the early Miocene (Agua Clara pull-apart basin developed east of the Guajira Peninsula
Formation, cycle C1), when the highest rate occurred (367 above the previous compressional setting (see Biju Duval
m/m.y). A high rate of 296 m/m.y. persisted for the rest et al., 1982). The locus of maximum subsidence within
of the Miocene. This area was uplifted during the late this pull-apart basin migrated eastward. The northern
Pliocene, so that cycle A rocks are absent. In summary, part of the Guajira Peninsula, however, developed
subsidence rates in the Urumaco trough were exceed- outside this extensional regime. During the early stages
ingly high, and they decreased and became more uni- of basin formation, maximum subsidence occurred in
form through time, until inversion of the depocenter in parts of offshore Aruba basin and in the northwestern
Recent time. part of the Falcón basin (Dabajuro platform), as well as in
the western Golfo de Venezuela.
Paraguaná Peninsula Deposition during late early Oligocene cycle C3 was
restricted to the offshore portion of the Guajira
Widespread sedimentation on the Paraguaná Peninsula, some areas of the Aruba region, and the
Peninsula began during cycle B (early middle Miocene). rapidly subsiding central part of the Falcón basin. Rapid
However, seismic interpretation suggests that sedimen- sedimentation began during cycle C2 (late Oligocene),
tation possibly started during the early Miocene in particularly in the Aruba region and the central Falcón
restricted northern areas. The early stages of Miocene basin. Rapid subsidence rates are calculated for the early
sedimentation record a moderate (93 m/m.y.) but highly Miocene (cycle C1) of the entire pull-apart system
variable sedimentation rate (Figures 18, 19). During the (100–140 m/m.y.), especially in the Urumaco trough and
rest of the Miocene (cycle B), sediments were deposited northeastern Falcón basin (over 300 m/m.y.), driven by
at a more moderate rate (120 m/m.y.), but again acceler- active growth along northwest-trending faults. High
ated during the Pliocene to 192 m/m.y. In conclusion, in rates (about 100–140 m/m.y.) were maintained through-
the Paraguaná Peninsula, sedimentation rates show a out the region during the middle and late Miocene (cycle
progressive increase through time. B). The Urumaco trough and northeastern Falcón basin
continued with higher rates (295 and 200 m/m.y., respec-
Offshore Northeastern Falcón Basin and tively).
La Vela Bay Because of erosion, data for Pliocene sedimentation
are not accurate for the onshore regions. The locus of
This area had two patterns of subsidence. The part of maximum Pliocene subsidence was located in the Golfo
La Vela Bay adjacent to the Paraguaná Peninsula (La Vela Triste area, where sediments were deposited at a rate of
platform) records lower rates than southeastern localities. 330 m/m.y.
Maximum subsidence apparently took place near the
present coast (Boca Ricoa) (Figure 18) during deposition
of cycle C (mean subsidence rate of 326 m/m.y.). As in CONCLUSIONS
the Urumaco trough area, subsidence rates remained
high, but decreased during Miocene and Pliocene time There has been considerable debate about the exact
(194 and 127 m/m.y., respectively) (Figure 18). location and geometry of the Caribbean–South American
The only data available for offshore southeastern plate boundary. Most recent work agrees that the
Falcón (Golfo Triste) are from the Cayo Sal-1 well. An boundary is located along the South Caribbean marginal
almost constant sedimentation rate of about 90 m/m.y. is fault and continues eastward into the El Pilar fault of
observed for the Eocene–late Miocene. This was followed northeastern Venezuela (Beck and Stephan, 1979;
by a sudden increase in the rate to 320 m/m.y. in the Stephan, 1985). The tectonic history of the Caribbean
Pliocene. chain has been summarized by Bellizzia (1972), Maresch
(1974), and Bellizzia et al. (1980), among others. During
Aruba the Jurassic, rifting occurred in the proto-Caribbean. An
island arc system collided obliquely with the continent in
Aruba subsidence rates are based on interpretation of Late Cretaceous–Oligocene time (Erlich and Barrett,
seismic data. Nine subsidence curves were constructed. 1990; Pindell, 1991), and a foredeep developed where the
Maximum rates occurred during deposition of cycle D crust was loaded. Eventually the allochthonous terranes
(late Eocene?), with a mean value of 226 m/m.y. in were thrust across Paleocene–Oligocene foredeep flysch
restricted depocenters. Subsidence rates remained more (Stephan, 1977, 1985; Beck, 1978).
or less constant during the late Oligocene–late Miocene, A major change in the tectonic regime took place in
with a mean value of 138–147 m/m.y. (Figures 18, 19). the Eocene as the Caribbean plate moved eastward with
Rates finally decreased to a mean of 38 m/m.y. during respect to South America (Stainforth, 1969; Bell, 1972;
the Pliocene (cycle A). In conclusion, subsidence rates in Maresch, 1974). East-west right-lateral transcurrent
Aruba were initially high, then decreased through time. faulting began in northern South America by late Eocene
time. This coincided with andesitic volcanism and
Comparison of Subsidence Rates subduction along the Lesser Antilles arc (e.g., Wester-
camp et al., 1985).
A regional right-lateral strike-slip setting developed East-west displacement between the two plates
during the late Eocene. As a result, a rapidly subsiding resulted in the formation of an extensive pull-apart zone
778 Macellari

During the Oligocene and Miocene, the allochthonous


thrust sheet was disrupted by stretching of the pull-apart
basin, and the axis of maximum subsidence migrated
eastward into the Urumaco trough and east of La Vela
Bay (Figure 20). Continued displacement along the Oca
fault resulted in transpressive deformation of older strata
and the formation of a thrust front that advanced
progressively northward. Thus, an ENE-WSW oriented
foreland basin began to form during this time.
Transpressive movement along the South Caribbean
marginal fault (Kellogg and Bonini, 1982) during the
Pliocene resulted in shallow underthrusting of the
Caribbean plate beneath the Curaçao Ridge. During this
time, the thrust front of the Falcón anticlinorium
migrated farther northward and extension along
northwest-southeast faults ceased. Active right-lateral
displacement along the Boconó fault began at this time
(Schubert and Sifontes, 1970). This displacement,
coupled with continued movement of the Oca fault
system, resulted in formation of a pull-apart depocenter
southeast of Falcón basin (Figure 20) (Schubert, 1984) and
thick Pliocene sedimentation.

Acknowledgments This study was funded by the Hocol


subsidiary of Tenneco Oil Co. I am grateful to D. Wimann
(British Gas, Ecuador) for his comments and support. T. Boesi,
C. Blanco (Maraven, Venezuela), and T. Subieta (Lagoven,
Venezuela) provided generous cooperation and insight during
my work in Venezuela. S. Schamel (Earth Sciences and
Resources Institute, University of South Carolina) offered
valuable advice during the earlier stages of this study.

REFERENCES CITED
Audemard, F., and I. J. Demena Arenas, 1985, Falcón oriental,
nueva interpretación estructural: VI Congreso Geológico
Venezolano, Memorias 4, p. 2317–2319.
Beck, C. M., 1978, Polyphasic Tertiary tectonics of the Interior
Figure 20—Evolutionary model for the southwestern Range in the central part of the western Caribbean Chain,
Caribbean pull-apart basin. Guarico State, northern Venezula: Geologie en Mijnbouw,
v. 57, p. 99-104.
Beck, C. M., and J. F. Stephan, 1979, Les grandes failles de la
marge Sud-Caraibe definissent elles la frontiere meri-
along the northern margin of South America, between donale d’une Plaque Caraibe?: 7th Reunion Annuelle des
the Curaçao Ridge and the San Sebastián–Oca fault Sciences de la Terre, Lyon, p. 237–238.
system (Figure 1). Right-lateral shearing caused exten- Beets, D. J., 1972, Lithology and stratigraphy of the Cretaceous
sional subsidence along northwest-trending normal and Danian succession of Curaçao: Publication of Natuur-
faults (Biju-Duval et al., 1982; Muessig, 1984). Basaltic weetenschapen Studiekring Suriname en Nederlandse
alkaline magmas were injected as a result of this crustal Antillen Uitgaven, no. 70, 153 p.
thinning (Muessig, 1984). It is believed that the Los Beets, D. J., H. J. MacGillavry, and G. Klaver, 1977, Geology of
Monjes and La Orchila islands, the Paraguaná Peninsula, the Cretaceous and early Tertiary of Bonaire: Guide to the
and the Falcón basin were almost juxtaposed before field excursions on Caraçao, Bonaire, and Aruba, Nether-
lands Antilles, Caribbean Geological Conference, VIII,
Oligocene extension (Muessig, 1978, 1984). Eocene Curaçao, Netherlands Antilles, p. 18–28.
isopachs (Figure 15) support a pull-apart model in which Bell, J. S., 1972, Global tectonics in the southern Caribbean
Curaçao and Bonaire were east of the Paraguaná area: GSA Memoir 132, p. 369–386.
Peninsula. During this time, the locus of maximum Bellizzia, G. A., 1972, Is the entire Caribbean mountain system
subsidence developed west of the allochthonous terrane of northern Venezuela allochthonous?: GSA Memoir 132,
(Figure 20). p. 363–368.
Cenozoic Sedimentation and Tectonics, Southwestern Caribbean Pull-Apart Basin 779

Bellizzia, G. A., R. Blanchet, C. Beck, and J. F. Stephan, 1980, Asociación Venezolana de Geologos Mineros y
La Chaine Caraibe, du Mesozoique à l’Actuel: tecto- Petroleros, Boletín Informativo, v. 10, p. 51–69.
genèse et modèle d’évolution géodynamique: 8th Halse, G. W., 1937, La estratigrafía del Occidente del Distrito
Reunion Annuelle des Sciences de la Terre, Marseille, p. Buchivacoa, Estado Falcón, Venezuela: Boletín de Geología
112–113. y Minería, Caracas, v. 1, p. 183–193.
Biju-Duval, D., A. Mascle, and P. Muzelec, 1979, Rapport Hunter, V. F., 1972, A middle Eocene flysch from east Falcón,
d’interprétation de la campagne seismique réflexion Venezuela: VI Caribbean Geological Conference,
Venézuela. I. Marge Nord Venezuela: Institut Francais du Margarita, Venezuela, Memoir, p. 126–130.
Petrole, Rapport geologique, no. 23650, 53 p. Hunter, V. F., 1974, The mid-Tertiary stratigraphic unit of the
Biju-Duval, D., A. Mascle, H. Rosales, and G. Young, 1982, southern Caribbean area, in P. Jung, ed., Contributions
Episutural Oligo-Miocene basins along the north dedicated to the geology and paleobiology in the
Venezuelan margin, in J. S. Watkins and C. L. Drake, eds., Caribbean and adjacent Areas—The Kugler Volume:
Studies in continental margin geology: AAPG Bulletin, v. Naturforschende Gesellschaft Basel, Verhandlungen, v. 84,
39, p. 347–358. p. 172–190.
Blow, W. H., 1959, Age correlation and biostratigraphy of Hunter, V. F., 1978, Foraminiferal correlation of Tertiary
the upper Tocuyo (San Lorenzo) and Pozon Formations, mollusc horizons of the southern Caribbean area: Geologie
eastern Falcón, Venezuela: Bulletin American Paleon- en Mijnbouw, v. 57, p. 193–203.
tology, v. 39, p. 67–252. Hunter, V. F., 1986, Tertiary event stratigraphy of the
Boesi, T., and D. Goddard, 1991, A new geologic model southern Caribbean borderland (abs.): 11th Caribbean
related to the distribution of hydrocarbon source rocks in Geological Conference, Barbados, p. 44.
the Falcón basin, northwestern Venezuela, in K. T. Biddle, Hunter, V. F., and Bartok, P., 1974, The age and correlation of
ed., Active margin basins: AAPG Memoir 52, p. 303–319. the Tertiary sediments of the Paraguaná Peninsula,
Bürgl, H., 1960, Geología de la Península de la Guajira: Venezuela: Asociación Venezolana de Geologos Mineros y
Boletín de Geología, Caracas, v. 6, p. 129–168. Petroleros, Boletín Informativo, v. 17, p. 143–159.
Case, J. E., T. L. Holcombe, and R. G. Martin, 1984, Map of Kellogg, J. N., and Bonini, W. E., 1982, Subduction of the
geologic provinces in the Caribbean region, in W. E. Caribbean plate and basement uplifts in the overriding
Bonini, R. B. Hargraves, and R. Shagam, eds., The South American plate: Tectonics, v. 1, p. 251–276.
Caribbean–South American plate boundary and regional Ladd, J. W., M. Truchan, M. Talwani, P. Stoffa, P. Buhl, R.
tectonics: GSA Memoir 162, p. 1–30. Houtz, A. Mauffret, and G. Westbrook, 1984, Seismic
Curet, E. A., 1992, Stratigraphy and evolution of the Tertiary reflection profiles across the southern margin of the
Aruba basin: Journal of Petroleum Geology, v. 15, Caribbean, in W. E. Bonini, R. B. Hargraves, and R.
p. 283–304. Shagam, eds., The Caribbean–South American plate
Diaz de Gamero, M. L., 1968, Paleontologia de la Formacion boundary and regional tectonics: GSA Memoir 162,
El Veral (Mioceno), Estado Falcón: GEOS, Universidad p. 153–159.
Central de Venezuela, Caracas, no. 17, p. 7–51. Liddle, R. A., 1946, The geology of Venezuela and Trinidad,
Diaz de Gamero, M. L., 1977a, Estratigrafía y micropaleon- 2nd edition: Ithaca, New York, Paleontological Research
tologia del Oligoceno y Mioceno inferior del centro de la Institute, 890 p.
Cuenca de Falcón, Venezuela: GEOS, Universidad MacDonald, W. D., 1964, Geology of the Serrania de Macuira
Central de Venezuela, Caracas, v. 22, p. 3–60. area, Guajira Peninsula, Colombia: Ph.D. dissertation,
Diaz de Gamero, M. L., 1977b, Revisión de las unidades Princeton University, Princeton, New Jersey, 167 p.
litoestratigráficas en Falcón Central en base a su Maresch, W. V., 1974, Plate tectonic origin of the Caribbean
contenido de foraminíferos planctónicos: V Congreso mountain system of northern South America: Discussion
Geológico Venezolano Memoria 1, p. 81–86. and proposal: GSA Bulletin, v. 85 , p. 669–682.
Diaz de Gamero, M. L., 1985, Estratigrafía de Falcón norori- Ministerio de Minas e Hidrocarburos, Venezuela, 1970, Léxico
ental: VI Congreso Geológico Venezolano Memoria 1, estratigráfico de Venezuela: Boletín de Geología, Caracas,
p. 454–502. Publicación Especial, 1, p. 1–728.
Duque-Caro, H., 1974, Los foraminíferos planctónicos y el Muessig, K. W., 1978, The central Falcón igneous suite,
Terciario de Colombia: Revista Española de Micropaleon- Venezuela: alkaline basaltic intrusions of
tologia, v. 7, p. 403–427. Oligocene–Miocene age: Geologie en Mijnbouw, v. 57,
Edgar, N. T., J. I. Ewing, and J. Hennion, 1971, Seismic p. 261–266.
refraction and reflection in the Caribbean Sea: AAPG Muessig, K. W., 1984, Structure and Cenozoic tectonics of the
Bulletin, v. 55, p. 833–870. Falcón basin, Venezuela, and adjacent areas, in W. E.
Erlich, R. N., and S. F. Barrett, 1990, Cenozoic plate tectonic Bonini, R. B. Hargraves, and R. Shagam, eds., The
history of the northern Venezuela-Trinidad area: Caribbean-South American plate boundary and regional
Tectonics, v. 9, p. 161–184. tectonics: GSA Memoir 162, p. 217-230.
Feo-Codecido, G., 1972, Breves ideas sobre la estructura de Payne, A. L., 1951, Cumarebo oil field, Falcón, Venezuela:
la Falla de Oca, Venezuela: IV Caribbean Geological AAPG Bulletin, v. 35, p. 1850–1878.
Conference, Margarita, Venezuela, Memoir, p. 184–190. Petzall, C., 1959, Estudio de una sección de la Formación
Gonzalez de Juana, C., 1937, General geology and strati- Caujarao en el anticlinal de La Vela, Estado Falcón:
graphy of Cumarebo area, State of Falcón: Boletín de Asociación Venezolana de Geologos Mineros y Petroleros,
Geología y Mineria, Caracas, v. 1, p. 187–208. Boletín Informativo, v. 2, p. 269–319.
Gonzalez de Juana, C., J. M. Iturralde de Azorena, and X. Pijpers, P. J., 1933, Geology and paleontology of Bonaire
Picard, 1980, Geología de Venezuela y de sus cuencas (Dutch West Indies): Ph.D. dissertation, University of
petrolíferas: Ediciones Foninves, Caracas, 2 vols., 1031 p. Utrecht, The Netherlands, 103 p.
Guevara, E. H., 1967, Contributions of the AVGMP Pindell, J. L., 1991, Geologic rationale for hydrocarbon explo-
Maracaibo basin Eocene nomenclature committee; VI, ration in the Caribbean and adjacent regions: Journal of
The Santa Rita, Jarillal, and La Victoria formations: Petroleum Geology, v. 14, p. 237–257.
780 Macellari

Renz, M. M., 1948, Stratigraphy and fauna of the Agua Salada Talwani, M., C. C. Windisch, P. L. Stoffa, P. Buhl, and R. E.
Group, State of Falcón, Venezuela: GSA Memoir 32, 219 p. Houtz, 1977, Multichannel seismic study in the
Renz, O., 1956, Cretaceous in western Venezuela and the Venezuelan basin and the Curaçao Ridge, in M. Talwani
Guajira (Colombia): 20th Session, International Geological and W. C. Pitman, eds., Island arcs, deep sea trenches, and
Congress, Mexico City, 13 p. back-arc basins: American Geophysical Union, Maurice
Renz, O., 1960, Remarks on the Barquisimeto trough: Ewing Series, no. 1, p. 83–98.
Asociación Venezolana de Geologos Mineros y Petroleros, Thomas, D. J., 1972a, El Eoceno de la Península de la Goajira:
Boletín Informativo 3, p. 155–162. IV Congreso Geológico Venezolano Memorias, v. 2,
Rodriguez, S. E., 1968, Estratigrafía y paleontologia del p. 951–962.
Mioceno en la Península de Paraguaná, Estado Falcón: Thomas, D. J., 1972b, The Tertiary geology and systematic
Asociación Venezolana de Geologos Mineros y Petroleros, paleontology (Phylum Mollusca) of the Guajira Peninsula,
Boletín Informativo, v. 2, p. 125–152. Colombia, South America: Ph.D. dissertation, State Univer-
Rollins, J. F., 1965, Stratigraphy and structure of the Guajira sity of New York at Binghamton, 147 p.
Peninsula, northwestern Venezuela and northeastern Tschanz, C. M., R. F. Marvin, B. J. Cruz, J. H. Mehnert, and G.
Colombia: University of Nebraska Studies (n.s.) 30, 102 p. T. Cebula, 1974, Geologic evolution of the Sierra Nevada
Schubert, C. , 1984, Basin formation along the Boconó-Morón- de Santa Marta, northeastern Colombia: GSA Bulletin, v.
El Pilar fault system, Venezuela: Journal of Geophysical 85, p. 271–284.
Research, v. 89, p. 5711–5718. Vail, P., and J. Hardenbol, 1980, Sea-level changes during the
Schubert, C., and R. S. Sifontes, 1970, Boconó fault, Tertiary: Oceanus, v. 22, p. 71–79.
Venezuelan Andes: Science, v. 175, p. 560–561. Vallenilla L., P., 1961, Estratigrafía de las formaciones
Senn, A., 1935, Die stratigraphische Verbreitung der Tertiaren Caujarao, La Vela, y Coro en sus localidades tipo, Estado
Orbitoiden, mit spezieller Beruchsichtigung ihres vorkom- Falcón: Asociación Venezolana de Geologos Mineros y
mens in Nord-Venezuela und Nord-Marokko: Eclogae Petroleros, Boletín Informativo, v. 4, p. 29–78.
Geologicae Helvetiae, v. 28, p. 51–113. Vasquez, E., 1975, Results of exploration in La Vela Bay:
Silver, E. A., J. E. Case, and H. J. Macgillavry, 1975, Geophys- Proceedings, IX World Petroleum Congress, Tokyo, v. 3,
ical study of the Venezuelan borderland: GSA Bulletin, v. p. 195–197.
86, p. 213–226. Westercamp, D., P. Andreieff, Ph. Bouyee, A. Mascle, and J. C.
Stainforth, R. M., 1962a, The upper Eocene of the Guajira Baubron, 1985, The Grenadines, southern Lesser Antilles,
Peninsula: Asociación Venezolana de Geologos Mineros y part I: stratigraphy and volcano-structural evolution, in A.
Petroleros, Boletín Informativo, v. 5, p. 229–230. Mascle, ed., Geodynamique des Caraibes: Paris, Editions
Stainforth, R. M., 1962b, Definition of some new stratigraphic Technip, p. 109–118.
units in western Venezuela: Las Pilas, Cocuiza, Vergel, El Wheeler, C. B., 1960, Estratigrafía del Oligoceno y Mioceno
Jebe, Tres Esquinas, and Nazareth: Asociación Venezolana inferior de Falcón occidental y nororiental: III Congreso
de Geologos Mineros y Petroleros, Boletín Informativo, v. Geológico Venezolano, Caracas, Memoria 1, Boletín
5, p. 279–282. Geología, Publicación Especial 7, p. 407–465.
Stainforth, R. M., 1969, The concept of sea floor spreading Wheeler, C. B., 1963, Oligocene and lower Miocene stratig-
applied to Venezuela: Asociación Venezolana de Geologos raphy of western and northeastern Falcón basin,
Mineros y Petroleros, Boletín Informativo, v.12, p. 257–274. Venezuela: AAPG Bulletin, v. 47, p. 35–68.
Stephan, J. F., 1977, El contacto cadena Caribe-Andes
Merideños entre Carora y El Tocuyo (Estado Lara): V
Congreso Geologico Venezolano, Caracas, Memorias, v.
11, p. 789–816. Author’s Mailing Address
Stephan, J. F., 1985, Andes et chaine Caraibe sur la transver-
sale de Barquisimeto (Venezuela): évolution géody- C. E. Macellari
namique, in A. Mascle, ed., Géodynamique des Caraibes: Nederlandse Aardolie Maatschappij B.V.
Paris, Editions Technip, p. 505–529. P.O. Box 28000
Sutton, F. A., 1946, Geology of the Maracaibo basin, 9400 HH Assen
Venezuela: AAPG Bulletin, v. 30, p. 1621–1741. The Netherlands
Index
Accretion, in Andes, 102 Andes, see also Eastern Cordillera Araucanian orogeny, 71-73, 394
Adrian Jara Formation, stratigraphy of, Aptian-Albian paleogeography of, Arches, of South America, 65
195 112-113 Architecture
Africa (southern), see also Pan-Africa and asthenosphere, 476-477 of Altiplano, 312-313, 320
basin evolution in, 5-44 backthrusting in, 475 of Gondwana, 8
data point analysis, 84 basement of, 271, 273-274 of South America, 65
Middle Devonian of, 257 basin-bounding features of, 243-244 of Venezuelan Coast Ranges crust,
Aguada Villanueva, seismic reflection bend in, see Andean bend 667-677
profile of, 395 blind thrusting in, 650-651 Arequipa massif, 233-234, 243
Aguaragüe field, cross section of, 552 Cambrian-Devonian evolution of, 214 Argentina
Aguaragüe range Carboniferous of, 233-237, 241 basement of, 271, 273-274
seismic reflection profile of, 534, 537 Cenozoic of, 118-119, 120-121, 122- basin evolution of, 251-265
structure of, 551-552 123, 225, 277-279 basin-bounding features of, 243-244
supersequence of, 255 of Colombia, 642-643 Cambrian-Ordovician of, 271-275
Algae, in organic matter, 414 compared to Canadian Cordillera, Carboniferous-Jurassic stratigraphy
Alluvial fans, of Sierra Pampeanas, 341- 653-654, 656 of, 257-261
356 Cretaceous rift system of, 325-337 Carboniferous-Permian of, 276, 285-
Altiplano Cretaceous paleogeography of, 110- 299
basin architecture of, 312-313, 320 111, 114-117 Cenozoic structures of, 277-279
basin evolution of, 305-322 cross sections of, 461, 465, 466, 467, correlation charts of, 134-135, 134-137,
Bouguer gravity of, 310, 313, 314 475, 477, 643 136-137, 138-139, 270
Carboniferous stratigraphy of, 233- crustal structure of, 639-642 Cretaceous unconformity of, 548
237 data point analysis of, 82 Cretaceous-Tertiary subsidence of,
correlation chart of, 321 Devonian of, 233-235, 240-241 276-277
geologic map of, 310-312 evolution of, 103-127 cross sections of, 153-158, 278
geologic setting of, 306-307, 512-514 geodynamics of, 244-245, 641-642 depocenters of, 275
geophysical data of, 308-309, 310-311, geologic setting of, 460-462, 512-514 gas fields of, 550
313-314, 322 hydrocarbon systems of, 104, 124-127 geologic maps of, 130-155, 272
hydrocarbons of, 307-308, 314-315 igneous events of, 242, 243-244, 271, geologic provinces of, 270
location map of, 512 273-274 glacial deposits of, 168-170, 173-174,
seismic reflection profiles of, 316-319 intermontane basins of, 597-611 174-176
stratigraphic columns of, 309, 311 Jurassic paleogeography of, 108-109 hydrocarbons of, 279-280, 365-366
stratigraphy of, 308, 312 lithosphere of, 461, 477, 642-643 igneous events of, 243-244, 271, 273-
structural inversion in, 321 location map of, 232, 648 274
structure of, 306, 315 new model for structure of, 652-653 paleogeography of, 383-400
and sub-Andean belt, 474 paleobiogeography of, 238-240 Permian of, 276
tectonostratigraphic chart of, 311 paleolatitudinal shift in, 231-246 petroleum geology of, 547-549
Alto Paraguay terrane, 39 palinspastic reconstruction of, 102- Phanerozoic history chart of, 270
Amazon basin, data point analysis of, 83 103 prerift geologic map of, 327
Ambo Group, 216, 235 Permian of, 233, 237-238, 241-243 regional map of, 404
Andean bend and pre-Andean extension, 490 rift system of, 307, 325-337
analogs of, 520 sandstone reservoirs of, 125-126 and sea level trends, 257
décollement levels in, 513 source rocks of, 124-125 sedimentation in, 177, 178
footwall morphology in, 515-516 strain of Tithonian rift of, 641-642 Silurian-Devonian stratigraphy of,
models of formation of, 518 structure of, 208-209, 514-518, 531 253-257, 275-276
orocline interpretation of, 518-521 subduction in, 643 structural inversion in, 341-356, 359-
paleogeography of, 518-519 tectonic evolution of, 269-281, 474- 366, 369-380
sedimentary thicknesses in, 517 477, 654 structure of, 549-554
structure of, 513-515 Tertiary compression in, 40 sub-Andean belt of, 549-554
Andean deformation, see also Andes Triassic-Jurassic paleogeography of, tectonic evolution of, 263, 269-281
in Boomerang Hills, 487-489 106-107 wildcat drilling in, 54-55
climax of, 74 Antarctica, data point analysis of, 84 Arica bend, lithospheric cross section of,
early, 73-74 Apon Formation 477
Ordovician, 212 isopach map of, 708 Aruba
seismic reflection profile of, 28 stratigraphy of, 707 geologic history of, 771-774

781
782 Index

interpreted seismic line of, 770 Jurassic, 70-72 tectonic map of, 232, 446
isopach map of, 772, 773, 775 late Paleozoic, 68-69 Triassic-Jurassic of, 219-221
stratigraphy of, 770-771 of Llanos basin, 660-663 wildcat drilling in, 54-55
subsidence rates of, 776, 777 of Malargüe basin, 371-373 Bolivia basin, summary of, 245
Asunción arch, cross section of, 85 in Mesozoic, 34-40 Bolivian Andes, data point analysis of,
Asthenosphere, and Andes, 476-477 Neocomian, 71-73 82
Atuel half-graben, cross section of, 377 of Neuquén basin, 383-400 Bolivian orocline, 512
Atuel-Valenciana half-graben of Paganzo basin, 288, 293-296 Bolsones, definition of, 68
cross section of, 375, 377-378 Phanerozoic, 66-67 Bonaire basin, geologic setting of, 759
geometry of, 374, 376 of Rumichaca basin, 607-608 Boomerang Hills
Ayacucho basin, stratigraphic column tectonic controls of, 5-44 Andean deformation in, 487-489
of, 602 Triassic, 69-70 cross section of, 447, 454, 456, 488
Azogues Formation, megaturbidites of, Basin resources, estimates of, 57-60 deformation summary of, 497
603 Bending, synorogenic, 512 fault map of, 487, 490
Azúcar Group Beni basin, stratigraphy of, 216 hydrocarbons of, 490-494
paleocurrents in, 621 Beni-Chaco plain, structural setting of, isopach map of, 498
stratigraphy of, 620-621 208 oil and gas window in, 456
Berta Formation, stratigraphy of, 195 oil fields of, 483
Backthrusting Blind thrusting, in Eastern Cordillera, petroleum system of, 447, 454
in Andes, 475 650-651 pre-Andean deformation of, 489-490
cross section of, 475 Bolivia sedimentary wedge of, 449
definition of, 650 Andean bend in, 511-521 seismic reflection profiles of, 485, 486,
Bahia Negra platform, hydrocarbons of, Andean deformation in, 481-498 491, 492, 493, 495, 496
199 basin-bounding features of, 243-244 stratigraphic column of, 483
Bajada Vidal trough, seismic reflection Cambrian-Caradoc of, 211-212 stratigraphy of, 483-486
profile of, 388 Carboniferous of, 32, 177, 215-218, structure of, 486-487, 488
Barda Colorado area, seismic reflection 233-237 Boomerang-Chiquitanas suture, 29
profile of, 393 Cenozoic of, 221-225 Bouguer gravity
Barinas-Apure basin Chaco basin of, 15 of Altiplano, 310, 313, 314
cross section of, 687, 693 correlation chart of, 132-133, 138-139 of Gondwana, 9
geologic setting of, 682 Cretaceous rift basin of, 305-322 Brasiliano cycle, 11-19, 32, 186-187
location map of, 682 cross sections of, 153-158, 222, 461 Brazil
Mesozoic-Cenozoic stratigraphy of, Devonian of, 213-215, 232-235 correlation charts of, 132-133, 134-135,
685-696 geochemical database of, 527 136-137
Paleozoic stratigraphy of, 682-685 geologic maps of, 130-155, 464, 482 geologic cross sections of, 153-158
Basement geologic setting of, 524 geologic maps of, 130-155
of Andes, 243-244 glacial deposits of, 174-176 glacial deposits of, 171-173, 176
of Argentina, 243-244, 271, 273-274 hydrocarbons of, 224-225, 451-457, wildcat drilling in, 54-55
basin forming, 66 526-531 Brazilian shield, igneous events of, 243
crustal evolution of, 8-10 Jurassic-Cretaceous of, 221 Bucaramanga fault, nature of, 651
extension in, 26 location map of, 460 Bulk strain, of Main sub-Andean thrust,
of Maracaibo basin, 705 Mesozoic paleogeography of, 218, 471-473
of Paganzo basin, 287 220, 223, 224 Buried thrust fronts, definition of, 650
of Patagonia, 405 oil correlation of, 527
shortening in Venezuelan Coast Ordovician of, 212, 213-215 Caipipendi
Ranges, 674 paleobiogeography of, 238-240 cross section of, 541
structural trends of, 9 Paleozoic paleogeography of, 210- geologic history chart of, 531
Basin evolution 211, 218 geologic map of, 539
and 2-D modeling, 728 Pennsylvanian-Triassic of, 218-219 HI/OI diagram of, 528
of Altiplano, 305-322 Permian of, 215 seismic reflection profile of, 540, 541
of Argentina, 251-265, 271-276 petroleum geology of, 224-225, 451- Calentura Formation, stratigraphy of,
block diagram of, 24 457 617
Carboniferous-Permian, 30-32 Phanerozoic stratigraphy of, 209-211, Campo Duran field
of Caribbean pull-apart basin, 778 226-227 cross section of, 553
of coastal Ecuador, 626-629 Silurian of, 213-215 seismic modeling of, 553
Cretaceous, 72-73 source rocks of, 448-450 Campo Duran-Madrejones range,
of Cuenca basin, 605-607 stratigraphic column of, 213, 215, 219, structure of, 552-554
early Paleozoic, 21-30, 68 221, 448 Canadian Cordillera, compared to
of Eastern Venezuelan basin, 743-745 stratigraphy of, 178, 254, 501-507 Andes, 653-654, 656
of intermontane basins, 604-608, 609, structural setting of, 208-209 Cancañiri Formation, stratigraphy of,
611 sub-Andean belt of, 445-457 211-212, 213
Index 783

Cangapi Formation, stratigraphy of, 259 Colombia of Bolivia, 153-158, 222, 461
Cape basin, evolution of, 21 Caribbean pull-apart basin of, 757- of Boomerang Hills, 488
Cape foldbelt, orogenesis of, 35 778 of Brazil, 153-158
Cape-Karoo basin Cordilleran crustal evolution of, 633- of Caipipendi, 541
as depocenter, 19-34 643, 647-657 of Campo Duran field, 553
subsidence curves of, 33 cross sections of, 635, 638, 642 of Capilla del Monte, 346
Capilla del Monte lithosphere in, 642-643 of Central thrust zone, 468
basement foliation of, 349 structure of, 651 of Chaco basin, 85, 189
block diagram of, 346 tectonic provinces of, 660 of Chaco-Paraná, 85-86
cross section of, 346 wildcat drilling in, 54-55 of Chile, 153-158
geology of, 345, 348 Colon Formation of coastal Ecuador, 627
Neogene thrusts of, 348-351 isopach map of, 708 of Cocuy basin, 640
normal faults of, 356 stratigraphy of, 707 of Colombia, 635, 638, 642
Caquiahuaca thrust, 464-467, 473 Colorado field, cross section of, 657 of Cuenca basin, 606
Caranavi anticline, structure of, 469 Competitor analysis, in petroleum of Cusiana field, 664
Carandaity basin exploration, 56-57 of Cuyo basin, 361, 362-364
hydrocarbons of, 199 Computer modeling, of Eastern of Devonian Andes, 244-245
stratigraphy of, 194 Venezuelan basin oil, 727-740 of Eastern Cordillera, 461, 547, 635,
Caribbean plate Copacabana Formation 638, 639, 640, 641, 642
margins of, 722, 757-778 reservoirs of, 433 of Eastern Venezuelan basin, 729,
tectonic provinces of, 758-759 as source rocks, 449, 450 744, 746, 747, 751
tectonic setting of, 700-702 setting of, 216, 427 of El Pilar fault, 673
Caribbean pull-apart basin, evolutionary stratigraphy of, 237, 238, 503 of Gondwana, 244
model for, 778 Cordillera, see Eastern Cordillera; of gravity-driven structures, 649
Castillo anticline, 409 Western Cordillera of Hollin Formation, 580, 586
Cauderalito limestone, structure at top Cordillera de la Costa, 759 of Ipaguazu field, 554
of, 769 Cordillera Oriental, see Oriente basin Jurassic-Recent, 86
Cayo Formation, stratigraphy of, 617-618 Cordillera Real, see Eastern Cordillera of Lomas de Olmedo basin, 337
Celiphus rallus-like algae, 414 Core photos of Llanos basin, 640
Central America, tectonic features of, 701 of Hollin Formation, 578 of Llanos foothills, 664
Central thrust zone of Manuripi X-1 well, 506 of Madrejones field, 554
cross section of, 468 of Pando X-1 well, 506 of Magdalena Valley, 652, 653, 657
structure of, 468-469 Coro high, stratigraphy of, 766 of Main sub-Andean thrust, 467, 468,
Cerro León Group, stratigraphy of, 190, Correlation charts 472
192-194 of Altiplano, 321 of Malargüe belt, 375-380
Cinco Picachos supersequence, 253-254 of Argentina, 270 of Mangan Formation, 607
Chronostratigraphy, see Stratigraphic of Chaco basin, 192 of Maracaibo basin, 687, 721
column, 406 of Cocinetas basin, 761 of Maturín basin, 747
Clay Pebble beds, stratigraphy of, 621- of Eastern Venezuelan basin, 745 of Merida arch, 709
622 of Falcon basin, 760 of Napo uplift, 580
Clinoforms, of Maracaibo basin, 714-716 of glaciated basins, 168 of Oriente basin, 655
Closed rift, definition of, 80, 81 of Madre de Dios basin, 427 of Paganzo basin, 294
Coastal Ecuador of Maracaibo basin, 703 of Palauco rift, 378
basin evolution of, 617-625, 626-629 of Marañon basin, 427 of Paraguay, 153-158, 197, 200
cross section of, 627 of Napo Formation, 583 of Paraná basin, 189
forearc basin of, 626-628 of Paganzo basin, 298 Permian-Triassic, 85
geologic setting of, 617 of South America, 132-139 of Pungarayacu area, 580
location map of, 616 of sub-Andean belt, 427, 431 of Ramos field, 551
paleocurrents of, 621, 624, 625 of Ucayali basin, 427 regional tectonostratigraphic, 23
paleogeography of, 618 of Venezuela, 684, 703, 760 of Rumichaca basin, 610
sedimentation rates of, 628 Critical taper, definition of, 473 of Serranía, 672, 673
stratigraphic column of, 618 Cross sections of Sierra de Pajarillo, 346
Coca-Payamino field, cross section of, of Aguaragüe field, 552 Silurian-Carboniferous, 85
586 of Andes, 461, 465, 466, 467, 475, 477, of southern South America, 153-156,
Cochabamba basin 643 157-158
cross section of, 515 of Argentina, 153-158, 278 of structural inversion, 321, 364
map view of, 515 of Asuñcion arch, 85 of sub-Andean belt, 447, 463, 533, 541,
structure of, 514, 515-516 Atuel-Valenciana half-graben, 375, 549
Cocinetas basin, stratigraphy of, 759-762 377-378 of triangle zone, 649
Cocuy basin, stratigraphy of, 640 of backthrusting, 475 of Tuichi syncline, 473
Codo del Senguerr anticline, 408 of Barinas-Apure basin, 687, 693 of Uruguay, 153-158
784 Index

of Venezuela, 754 stratigraphic column of, 188, 194, 196 of southern Africa, 84
of Venezuelan Coastal Ranges, 672, stratigraphy of, 216-218 of West Texas Permian basin, 83
673 and sub-Andean belt deformation, Delamination, by tectonic wedge, 551
of worldwide foldbelts, 470-471 535-536 Depocenters, of Paleozoic, 19-34
Crude oil, resources of Latin America, 54 subbasins of, 186, 196, 199 Deseado massif, definition of, 64
Crust Chaco Formation, stratigraphy of, 198 Diaguita orogeny, 276-277
of Andes, 639-642 Chacopampeana Plain, geologic setting Diamante River area, cross section of,
of Venezuelan Coast Ranges, 667-677 of, 252-253 377
Cuenca basin Chaco-Paraná basin Diamictites
cross section of, 606 glacial deposits of, 173-174 of Bolivian Andes, 234-236
evolution of, 605-607 sedimentation in, 177 of Itarare Group, 173
megaturbidites of, 603 Chacopata-Uverito transect of Tarija Formation, 179
paleogeography of, 606 location map of, 728 Dorsal de Huincal, see Huincal dorsal
stratigraphic column of, 601 oil maturity along, 753 Dry holes, percentage of, 55
structural map of, 608 Chaco-Tarija basin
tectonic setting of, 605 depositional setting of, 180 Eastern Cordillera
Cuevo supersequence, stratigraphy of, glacial deposits of, 174-176 angular unconformity of, 260
259 hydrocarbons in, 177-178 blind thrusting in, 650-651
Cumana Formation, lithologies of, 235 sedimentation of, 19-34, 178 compared to Canadian Cordillera,
Curaçao Ridge, setting of, 758 stratigraphic columns of, 178 653-654, 656
Curupaity subbasin, hydrocarbons of, Chañic belt, tectonic evolution of, 276 cross sections of, 461, 547, 635, 638,
199 Chañic orogeny, 89, 425-426 639, 640, 641, 642
Cusiana field Chapiza Formation, stratigraphy of, 562- crustal configuration of, 639-642
cross section of, 664 563 foreland basin of, 635-636
seismic reflection profile of, 637 Charagüe, seismic reflection profile of, geologic setting of, 252-253, 512-514,
Cuyo basin 537, 538 561
cross section of, 361, 362-364 Chenque-Challao anticline, 410 gravity-driven structures of, 648-649,
hydrocarbons of, 365-366 Chichibacoa basin 657
inversion in, 359-366, 364 offshore wells of, 762 hydrocarbons of, 654-656
seismic reflection profile of, 364 stratigraphy of, 762-763 location map of, 512
source rocks of, 365-366 Chile new model for structure of, 652-653
stratigraphic column of, 40 basin-bounding features of, 243-244 palinspastic restoration of, 639
stratigraphy of, 361, 363 correlation charts of, 132-133, 134-135, polyphase rifting in, 638
structure of, 360, 361-362 136-137 rift tectonics of, 639-639
tectonic evolution of, 362-363 Devonian stratigraphy of, 232-235 Silurian-Devonian unconformity in,
Cuyo Group geologic cross sections of, 153-158 255
isopach map of, 374 geologic maps of, 130-155 snip restoration of, 476
paleogeography of, 387 igneous events of, 244 stratigraphy of, 258, 259
stratigraphy of, 388-390 Permian paleogeography of, 242 structure of, 277, 461, 641, 642
transfer faults of, 399 regional map of, 404 tectonics of, 546, 634-636
rift basins in, 342 Eastern Venezuelan basin
Chacay Melehue half-graben, 375, 376 Chiquitanas-Boomerang suture, 18-19 computer modeling of, 727-740
Chacay-Lotena Group, isopach map of, Chongón-Colonche Cordillera correlation chart of, 745
375 stratigraphic column of, 620, 622 cross sections of, 729, 744, 746, 747,
Chaco basin tectonostratigraphic evolution of, 751
Carboniferous of, 177 619-623 geodynamic evolution of, 743-745
Cenozoic of, 189-190, 195-198 transgressive deposits of, 622, 623 HI/OI of, 752
cross sections of, 85, 189 Chonta Formation, oil-source correlation hydrocarbon saturation in, 735, 738,
data point analysis of, 82 of, 432 739
Escarpment Formation of, 32 Chota basin, stratigraphic column of, 601 hydrocarbons of, 727-740, 749-755
evolution of, 29 Chubut Group, 406 hydrodynamic model of, 750-751
geologic map of, 187, 482 lithofacies of, 731
geologic setting of, 546 Dahlstromian, definition of, 634 location map of, 728, 743
geothermal gradient of, 199 Data point analysis megasequences of, 743-745
hydrocarbon potential of, 198-201 of Amazon basin, 83 Miocene facies of, 736
isopach map of, 193 of Antarctica, 84 oil fields of, 742
Mesozoic of, 189, 195-196 of Bolivian Andes, 82 petroleum geology of, 749-755
Paleozoic of, 187-189, 190-195 of Chaco basin, 82 pressure zones in, 732-733
pre-Andes deformation of, 498 of Huallaga basin, 83 reservoirs of, 749-750
sedimentary episodes of, 186 of Paraná basin, 82 seismic reflection profiles of, 744, 747,
seismic reflection profile of, 15, 191 of Sierras Chiquitanas, 80 748, 749, 750
Index 785

source rocks of, 730, 737, 752 stratigraphy of, 763-768 of Malargüe belt, 370
stratigraphic columns of, 730, 745 subsidence rates of, 776 of Neuquén basin, 370
stratigraphy of, 730, 731, 735, 743-745 well correlation in, 768 of Occidental, 352
structure of, 746-748 Fan, definition of, 176 Ordovician, 130, 133, 141
tectonic setting of, 742 Fast flexure, 80, 81 of Paraguay, 187
vitrinite reflectance in, 734, 736, 737, Fault-bend fold, 487, 489 Permian, 139-142, 146, 147, 148
738 Field discoveries of Rumichaca basin, 610
water flow in, 734 major, 56-57 of Saldan Formation, 348
Eastern Venezuelan ranges, see also from wildcat drilling, 55 of Serranía, 670
Venezuelan Coast Ranges Field size of Sierra Chicas, 345
geodynamic modeling of, 675-677 95th percentile of, 59 of Sierra de Los Condores, 347
geologic setting of, 668-672 distribution of, 59-60 of Sierras de Cordoba, 343
subduction in, 675-676 in Latin America, 58-59 Silurian, 133, 135, 142
tectonostratigraphy of, 670-671 population statistics of, 58-59 of sub-Andean belt, 464, 469, 525
terranes of, 670-673 worldwide, 58-59 Tertiary, 150-153, 154, 155
Economic basement, see Basement Field size analysis, definition of, 57 Triassic, 143-144, 149, 150
Ecuador, see also Coastal Ecuador Flexural load model, of Andean crust, Vendian-Cambrian, 130, 140
coastal, 615-629 639-641 Ghanzi thrusts, seismic reflection profile
Cretaceous paleogeography of, 582- Flexure of, 12-13
585 fast, 80, 81 Glacially influenced, definition of, 167
depositional systems of, 575-582 slow, 81 Glaciated basins
geologic setting of, 574-575 Foldbelts, cross section of worldwide, correlation chart of, 168
intermontane basins of, 597-611 470 marine sediments in, 165-176
location map of, 574 Footwall morphology, in Andean bend, Glaciation
morphostructural map of, 598 515-516 of Gondwana, 31, 169
reservoirs of, 585-590 Forced fold, definition of, 489 and hydrocarbon reserves, 165-180
seismic expression of structures of, Forearc basin, of coastal Ecuador, 626- Middle Silurian, 88
559-570 628 sediments of, 165-176
stratigraphic column of, 601 Foredeep fill, of sub-Andean belt, 462 Golfo Tristel, stratigraphy of, 770
stratigraphy of, 575-582 Foreland basin, of Maracaibo basin, 717- Gondwana
tectonic map of, 560 719, 720, 721 basin evolution of, 5-44
tectonic setting of, 605 Foreland basins basin map of, 7, 9
wildcat drilling in, 54-55 of Eastern Cordillera, 635-636 Carboniferous paleogeography of,
El Furrial structure, 750 evolution of, 73-74 90-93, 169
El Pilar fault glacial deposits of, 167, 168-170 crustal evolution of, 8-10
cross section of, 673 hydrocarbons in, 176-177 Devonian of, 88-90, 244
and deep structure, 676 of Maracaibo basin, 710-719 evolution of, 208
geologic setting of, 668-672 Paleozoic of Andes, 231-246 glaciation of, 31, 88, 169, 235
Ene Formation, oil-source correlation of, of Peru, 424, 425 Ordovician of, 86-87
428 Paleozoic of, 212, 218
Entre Lomas trough, seismic reflection Gacela field, stratigraphic correlation of, Permian of, 92-95
profile of, 389 587 Phanerozoic evolution of, 66-67
Escarpment Formation, 32, 297 Gariep belt, 11 reconstruction map of, 81
Estancia Vieja area, seismic reflection Gas fields, see also Oil fields Silurian of, 88-89
profile of, 396 of Argentina, 550 structural trends of, 9
Etendeka magnetism, 39 largest in Latin America, 57 tectonic subsidence analysis of, 79-96
Eva-Eva thrust, 464-467 GENEX software, 450-451, 751 Triassic of, 95-96
Exploration efficiency Geochemistry Gravity-driven structures
definition of, 57-58 of organic matter, 414 cross section of, 649
for Maracaibo basin, 58 of sub-Andean belt, 437, 526-531 of Eastern Cordillera, 648-649, 657
for Maturin basin, 58 Geologic maps geometry of, 648-649
for Venezuela, 58 of Altiplano, 310-312 of Magdalena Valley, 649-650, 657
Explosive source data, for Salta rift, 339 of Argentina, 272 in Tien Shan mountains, 649
Extensional arc, definition of, 103 of Bolivia, 464 Guajira Peninsula
Carboniferous, 137, 139, 145 Cenozoic basins of, 761
Faja Petrolifera, 748 of Chaco basin, 187 geologic history of, 771-774
Falcón basin Cretaceous, 147-150, 152, 153 geologic setting of, 759
geologic setting of, 759 Devonian, 135, 137, 143, 144 isopach map of, 772, 773, 775
interpreted seismic line of, 763 of Inicua quadrangle, 469 stratigraphic column of, 760
paleogeography of, 764, 765, 767 Jurassic, 145-147, 151 stratigraphy of, 759-763
stratigraphic column of, 760 of Llanos foothills, 663 subsidence rates of, 774-776
786 Index

Guandacol supersequence, 288-290 in Latin America, 53-61


Guapore shield of Marañon basin, 424, 433-440 La Cruz Conglomerate, 354
stratigraphy of, 192 maturation of, 126-127, 415 La Luna Formation
structural setting of, 208 of Paganzo basin, 296 isopach map of, 708
Guasare Formation of Paraguay, 198 stratigraphy of, 707
isopach map of, 711, 713 of Patagonia, 413-416 La Paz Formation, stratigraphy of, 190-
stratigraphy of, 711-712 of Pirital area, 754 192
Guayaquil area of San Bernardo belt, 413-416 La Quinta Formation, Jurassic of, 706
stratigraphic column of, 618 of sub-Andean belt, 424, 433-440, 451- La Valenciana anticline, cross section of,
tectonostratigraphic evolution of, 457, 547-549 378
617-619 TEMISPACK used for, 728 La Valenciana half-graben, 373, see also
Guayaquil Formation, stratigraphy of, traps of, 415-416 Atuel-Valenciana half-graben
619 of Ucayali basin, 424, 433-440 La Vela Bay
Guayuta Group, as source rocks, 737 stratigraphy of, 769-770
Guyana shield, 673 Igneous events, of Andes, 243-244, 271, subsidence rates of, 777
273-274 wells in, 769
Hiatus, Carboniferous of Bolivia, 237 Imbrication, in sub-Andean belt, 464- Labrador Sea, channels in, 180
Historical field size analysis, definition 466, 471, 474 Lake Maracaibo, see Maracaibo basin
of, 57 Inca movements, 276-277 Las Breñas Formation, 328
Hollin Formation, see also Main Hollin Inicua quadrangle, geologic map of, 469 Las Heras Group, 405
Formation; Upper Hollin Intermontane basins Las Lenas anticline, cross section of, 377
Formation age of sedimentary rocks of, 599 Las Pavas supersequence
core photos of, 578 age of volcanic rocks of, 598-599 logs of, 256
Cretaceous paleogeography of, 584 basin evolution of, 609, 611 stratigraphy of, 254-255, 256
cross section of, 580, 586 definition of, 598 Las Penas Formation, stratigraphy of,
depositional systems of, 575-580 of Ecuador, 609, 611 258-259
isopach map of, 576 evolution of, 604-608 Las Piedras Formation, 730
logs of, 577, 580, 581 lithostratigraphy of, 600-604 Lateral ramp, geometry of, 654
petrography of, 588 location of Ecuadorian, 599 Latin America, see also South America
petrophysical properties of, 589-590 location of Peruvian, 600 hydrocarbons in, 53-61
reservoirs of, 585-590 Neogene of Andes, 597-611 petroleum basin map of, 54
roadcut of, 581 sequence stratigraphy of, 602-604 undiscovered resources of, 60
stratigraphic column of, 576 Intracratonic basins Laurussia, Carboniferous paleogeog-
stratigraphy of, 575-580 glacial deposits of, 171-176 raphy of, 169
wettability of, 590 hydrocarbons in, 177-180 Leeward Antilles terrane, 758-759
Huallaga basin, data point analysis of, 83 Intracratonic subsidence, definition of, Lewis thrust sheet, compared to Andes,
Huarina belt, stratigraphy of, 216 81 653-654, 656
Huincul dorsal Inversion, see Structural inversion Lias, 405
interpretation of, 394 Ipaguazu field, cross section of, 554 Line 25106-28, seismic reflection profile
seismic reflection profiles of, 38, 393, Ipaguazu Formation, stratigraphy of, of, 317
395, 396 259 Line 2585-28, seismic reflection profile
structural inversion in, 391-396 Isiboro cross section of, 319
tectonic evolution of, 384, 386 maturity in, 455 Line 2586-28, seismic reflection profile
Hydrocarbons oil and gas window of, 454 of, 318
of Altiplano, 307-308, 314-315 petroleum system of, 453-454 Line 2587-28, seismic reflection profile
and Andean development, 104, 124- Isiboro thrust, 465 of, 316
127 Isla del Sol, 235 Line 3272-26, seismic reflection profile
of Argentina, 265, 279-280, 365-366 Itapucumí Group, stratigraphy of, 190 of, 25
of Bolivia, 224-225, 451-457 Itararé Group Line 4032-24, seismic reflection profile
of Boomerang Hills, 490-494 depositional model of, 176 of, 26
of Chaco basin, 198-201 logs of, 172, 173, 174 Line 4196
of Cuyo basin, 365-366 moldic porosity in, 175 seismic reflection profiles of, 327, 329,
of Eastern Cordillera, 654-656 sedimentary columns of, 173 330-331, 334-336
of Eastern Venezuelan basin, 727-740, Izozog arch, seismic reflection profile of, strength attribute of, 328
749-755 27 Line 4542-20, seismic reflection profile
effect of faulting on, 365-366 of, 28
field discoveries of, 55-56 Jurásico cycle Line 5008-21, seismic reflection profile
in foreland basins, 176-177 block diagram of, 376 of, 27
in glacially influenced sediments, isopach map of, 373 Lisure Formation, stratigraphy of, 707
165-180 Lithosphere, in Andes, 461, 477, 642-643
in intracratonic basins, 177-180 Karoo basin, evolution of, 31 Lomas de Olmedo basin
Index 787

cross section of, 337 Magmatism, Triassic-Jurassic, 35-36 passive margin sedimentation of,
field parameters of, 340 Magnetism, of Paraná basin, 39 706-709
isopach map of, 326 Main Andean thrust rift sedimentation of, 705-706
seismic reflection profile of, 264 stages of thrust dynamics, 473-474 sandstone petrography of, 714
stratigraphic column of, 40 tectonics of, 470-471 seismic reflection profiles of, 701, 715-
structure of, 333 Main Hollin Formation, see also Hollin 717
Lonco Trapial, 405 Formation seismic stratigraphy of, 702-704
Longhorn basin, geometry of, 638-639 core permeability of, 595 stratigraphic columns for, 703, 704
Los Blancos anticline, cross section of, deposition systems of, 576-577 stratigraphy of, 682-696, 705-707, 711-
377 Mohr-Coulomb failure criteria of, 596 713
Los Monos Formation petrophysical properties of, 594 subsidence curves of, 709, 710, 715,
HO/OI diagram of, 529 porosity-permeability of, 593 719
as source rocks, 528-529, 547 Main sub-Andean thrust subsurface seismic map units of, 702-
stratigraphic column of, 529 balancing bulk strain in, 471-473 704
Los Roques basin, 758 cross section of, 467, 468, 472 well data of, 701, 702
Lotena Group, paleogeography of, 391 seismic reflection profile of, 472 Marañon basin
Lower Patquia–De la Cuesta superse- stages of thrust dynamics of, 473-474 correlation chart of, 427
quence, stratigraphy of, 291-292 tectonics of, 470-471 exploration wells of, 442
Main thrust, definition of, 469 hydrocarbons of, 424, 433-440
Llallagua Formation, stratigraphy of, 213 Main thrust tectonics, definition of, 469 oil fields of, 443
Llanos basin Malargüe basin, evolution of, 371-373 seismic reflection profile of, 426
basin evolution of, 660-663 Malargüe belt setting of, 425-426
chronostratigraphic summary of, 661 cross sections of, 375-380 Massifs, of South America, 65
cross section of, 640 geologic map of, 370 Maturation, mechanisms of, 126-127
logs of, 662 structural inversion in, 369-380 Maturín basin
stratigraphic column of, 662 Malvinas basin, evolution of, 71-73 cross section of, 672, 673, 747
stratigraphy of, 660-663 Malvinokaffric fauna, 239 exploration efficiency for, 58
Llanos foothills Manabí area, tectonostratigraphic field size distribution in, 59
chronostratigraphic summary of, 661 evolution of, 619 hydrodynamic system of, 751
cross section of, 664 Mandeyapecua thrust, interpretation of, tectonostratigraphy of, 671-672
geologic map of, 663 539 Megasequences
logs of, 662 Mandiyutí Group definition of, 211
stratigraphic column of, 662 evolution of, 297-298 of Eastern Venezuelan basin, 743-745
stratigraphy of, 660-663 setting of, 216-217 of Maracaibo basin, 703-704, 707, 711-
structure of, 663-664 stratigraphy of, 258-259 713
Lliquimuni cross section Mangán Formation, 607 Mendoza area, structural inversion in,
maturity in, 453 Manuripi X-1 well 369-380
oil and gas windows of, 451 core photos of, 506 Merida arch
petroleum system of, 447, 451-453 depositional systems of, 504-507 cross section of, 709
depth/dates of, 503 Jurassic rifts of, 705
Macarao Formation, stratigraphy of, 759 lithofacies of, 506 Mesa Formation, 730
Macharetí Group location of, 502, 503 Mesoproterozoic, crustal evolution in, 8-
evolution of, 297-298 stratigraphy of, 502-504 10
setting of, 216-217 Maraca Formation Mesosaurus, 31
stratigraphy of, 257-258 isopach map of, 708 Miraflores syncline, stratigraphic
McKenzie basin model, 638 stratigraphy of, 707 column of, 221
Macuma Formation, stratigraphy of, 561 Maracaibo basin Misoá Formation
Madre de Dios basin basement of, 705 clinoforms of, 715, 716
correlation chart of, 427 clinoforms of, 714-715 isopach map of, 711, 713
depositional systems of, 504-507 correlation chart for, 703 progradation in, 717
exploration wells of, 442 cross sections of, 687, 721 seismic reflection profile of, 715, 716
setting of, 426 depositional environments of, 704 stratigraphic column of, 693
stratigraphy of, 501-507 exploration efficiency for, 58 stratigraphy of, 712-713
Madrejones field, cross section of, 554 fault map of, 718 Morichito basin
Magallanes basin, evolution of, 71-73 foreland basins of, 710-719, 720, 721 cross section of, 672, 673
Magdalena Valley fossil assemblages of, 704 seismic reflection profile of, 671
cross section of, 652, 653, 657 geologic/tectonic setting of, 682, 700- Mulichinco Formation, paleogeography
gravity-driven structures of, 649-650, 702 of, 397
657 isopach maps of, 708, 711, 712, 713 Multinational oil companies (MNCs),
low-angle normal faults of, 656 location map of, 682 definition of, 56
structure of, 651 paleocurrents of, 714-715
788 Index

Nama Group Occidental evolution of, 31, 288, 293-296


basin fill of, 16 Cretaceous stratigraphy of, 352-353 hydrocarbon potential of, 296
deposition of, 15-16 geologic map of, 352 paleogeography of, 296-298
outcrops of, 16 Neogene inversion of, 353-356 seismic reflection profile of, 289
seismic reflection profiles of, 12, 14 Neogene structure of, 351-352 stratigraphic column of, 287
Namora basin, stratigraphic column of, stratigraphic column of, 354 stratigraphy of, 288-293
602 Ocloyic foldbelt subsidence curves of, 33
Napo, reservoir petrography of, 586-589 regional framework of, 273 supersequences of, 288-296
Napo Formation tectonic evolution of, 275 tectonic model for, 294, 295
correlation chart of, 583 Oil fields Paganzo-Maliman basin
Cretaceous paleogeography of, 585 of Argentina, 265 glacial deposits of, 168-170
depositional systems of, 580-582 of Boomerang Hills, 483 stratigraphic framework of, 169
idealized depositional package of, of Cuyo basin, 362-363, 365-366 Palauco half-graben, geometry of, 374,
582 of Eastern Venezuelan basin, 742 376
logs of, 581 largest in Latin America, 57 Palauco rift, cross section of, 378
petrophysical properties of, 589-590 of Marañon basin, 443 Paleobiogeography
stratigraphy of, 580-582 of Neuquén basin, 386 of Carboniferous Andes, 239-240
T and U sandstone reservoirs of, 581- of Peru, 424 of Devonian Andes, 238-239
582 of San Bernardo belt, 405 of Permian Andes, 240
transgressive-regressive cycle of, 583 sizes of, 59-60 Paleogeography
Napo uplift, cross section of, 580 of Ucayali basin, 443 Albian, 112-113
National oil companies (NOCs), defini- Oil and gas, see Hydrocarbons of Andean bend, 518-519
tion of, 56 Open rift, definition of, 80, 81 of Andes, 240-243
Nazca plate, 277 Organic matter, geochemical profile of, Aptian, 112-113
Neoproterozoic 414 of Argentina, 383-400
basins of, 7 Oriente basin of Bolivia, 210-211, 218, 220, 240-243
extension in, 12 Cretaceous paleogeography of, 582- Cambrian, 20, 42, 67
outcrop belts of, 10 585 Campanian, 116-117
of Pan-Africa, 17 cross section of, 655 Carboniferous, 42, 67, 90-93, 169, 241,
tectonic reconstruction of, 18 depositional systems of, 575-582 297
Netherland Antilles, 758-759 evolution of, 655, 636 Cenomanian, 114-115
Neuquén basin geologic setting of, 561, 574-575 Cenozoic, 224
early rifting in, 386 igneous events of, 243 of coastal Ecuador, 618, 625
evolution of, 71-73 location map of, 574 construction of maps of, 103
geologic map of, 370 logs of, 581 Cretaceous, 37, 43, 71, 72, 73, 223, 224,
geologic setting of, 386 reservoirs of, 585-590 263, 344-348, 397, 399, 584, 585,
Huincal arch in, 38 seismic reflection profiles of, 563, 564, 688-692
isopach map of, 390 565, 566, 567, 568, 569, 636, 637 of Cuenca basin, 606
Jurassic rift of, 373-375 stratigraphic columns of, 562, 575, of Cuyo Group, 387
location map of, 384 576 Devonian, 24, 42, 67, 88-90, 234, 240-
oil fields of, 386 stratigraphy of, 561-565, 575-582 241
paleogeography of, 387, 391 structure of, 565-569 of Ecuador, 582-585
seismic reflection profiles of, 388, 389, tectonic map of, 560 Eocene, 118-119, 694, 695
393, 395, 396 Orinoco oil belt, cross section of, 672-673, of Falcón basin, 764, 765, 767
stratigraphic column of, 372 729 of Hollin Formation, 584
stratigraphy of, 385, 388-391 Oritupano area, porosity vs. depth in, Jurassic, 34, 37, 43, 70, 71, 106-107,
structural inversion in, 369-380, 391- 733 108-109, 220, 373, 387, 391, 392, 399
393 Orocline late Paleozoic, 69
structural style of, 386-388 Andean bend as, 518-521 of Lotena Group, 391
subsidence in, 391-394 definition of, 512 Maastrichtian, 116-117
Tertiary folding of, 375-380 Orogenesis of Mesozoic Andes, 118-124
Neuquén Group Andean, 696; see also Andes Middle Silurian, 88-89
cross section of, 398 Araucanian, 71-73 Miocene, 120-121, 122-123, 696, 765,
isopach map of, 399 Permian-Triassic, 34-35 767
paleogeography of, 397 Tertiary, 40 of Mulichinco Formation, 397
Tertiary orogenesis of, 397-398 Vendian, 15-19 of Napo Formation, 585
Normal fault, horizontal offset from, 653 Neocomian, 110-111
North Patagonia massif, definition of, 64 Paganzo basin, 276 of Neuquén basin, 387, 391
“Northwestern basin,” 252 basement of, 287 of Neuquén Group, 397
correlation chart for, 298 Oligocene, 120-121, 695, 764
Oca fault, well correlation across, 768 cross section of, 294 Ordovician, 20, 42, 67, 86-87
Index 789

of Oriente basin, 582-585 Paraná basin, 11 of Boomerang Hills, 447, 454


Paleocene, 118-119, 692, 694 cross sections of, 85, 189 of Isiboro cross section, 447, 453-454
of Paleozoic Andes, 106-117 data point analysis of, 82 of Lliquimuni cross section, 447, 451-
Permian, 30, 42, 67, 70, 92-95, 218, as depocenter, 19-34 453
241-243 glacial deposits of, 171-173 of sub-Andean belt, 452-455
Pleistocene, 696 hydrocarbons of, 178-180 Phanerozoic
Recent, 122-123 isopachs, 39, 193 basin evolution in, 5-44, 66-67
Santonian, 114-115 logs of, 172 stratigraphic column of, 21
of Sierra Chicas, 344-348 magnetism of, 39 Pilar subbasin, hydrocarbons of, 199
Silurian, 67 outcrop map of, 175 Piñon Formation, stratigraphy of, 617
of southern hemisphere, 86-96 subsidence curves of, 33 Pirgua Subgroup, stratigraphy of, 330-
Tertiary, 73, 263 Parasequence, of Las Pavas, 255, 256 332
of Tordillo Formation, 392 Passage beds, stratigraphy of, 621 Pirital area, hydrocarbons of, 754
Triassic, 34, 43, 67, 95-96, 106-107, 218, Passive roof duplexes, definition of, 650 Pirital high
220, 387 Patagonia cross section of, 672, 673, 729
Vendian, 42 basement of, 405 seismic reflection profile of, 671
of western Venezuela, 683, 688-692, hydrocarbons of, 413-416 Pirity subbasin
694-696 tectonic evolution of, 403-413 hydrocarbons of, 199
Palinspastic reconstruction, of Andes, Patquía–De la Cuesta supersequence, seismic reflection profile of, 191
102-103 stratigraphy of, 291-292 Polyphase rifting, in Eastern Cordillera,
Palmar de las Islas Group, stratigraphy Paují Formation 638
of, 195 isopach map of, 711, 713 Portete basin, 762-763
Palo Santo Formation stratigraphy of, 713 Pre-Andean deformation
seismic contours of, 197 Pelado structure, maturity of, 452 of Boomerang Hills, 489-490
stratigraphy of, 195-196 Perija Andes, compared to Canadian of Chaco basin, 498
Pampean Ranges, 68, 271 Cordillera, 653-654, 656 Precambrian belts
igneous events of, 243-244 Peru of Argentina, 271, 273-274
location map of, 343 basin-bounding features of, 243-244 of South America, 286
stratigraphic column of, 344 Carboniferous stratigraphy of, 233- Pucara Group, isopach map of, 429
structural inversion in, 341-356 237 Pull-apart basin, of SW Caribbean, 757-
Pan-Africa, tectonic framework of, 11-19 cross section of, 425 778
Pando X-1 well Devonian stratigraphy of, 232-235 Pumbuiza Formation, stratigraphy of,
core photos of, 506 foreland basins of, 424, 425 561
depositional systems of, 504-507 igneous events of, 243 Puna arch, evolution of, 29
depth/dates of, 503 intermontane basins of, 597-611 Puna-Pampeanas arch, 29, 30, 33
lithofacies of, 506 morphostructural map of, 598 Pungarayacu area
location of, 502, 503 oil fields of, 424 cross section of, 580
stratigraphy of, 502-504 Permian paleogeography of, 242 depositional systems of, 579-580
Pangea, late Paleozoic reconstruction of, petroleum geology of, 423-444 Punilla thrust
212, 226 stratigraphic column of, 602 basement foliation of, 349
Pan-Gondwana, tectonic framework of, wildcat drilling in, 54-55 slip analysis of, 349, 350
11-19 Petroleum basins; see also Oil fields Punta Ancón Formation, stratigraphy of,
Paraguaná Peninsula of Latin America, 54 623-624
interpreted seismic line of, 769 of South America, 54, 63-74
isopach map of, 772, 773, 775 Petroleum exploration, see also Hydro- Quechua orogeny, 276-277
stratigraphic column of, 760 carbons
stratigraphy of, 768-769 competitor analysis of, 56-57 Ramos field, cross section of, 551
subsidence rates of, 776, 777 Petroleum geology Ramos range, structure of, 550-551
Paraguaná reflector, 769 of Argentina, 547-549 Ramp, definition of, 487
Paraguay of Bolivia, 224-225 Reflection seismic profiles, see Seismic
correlation charts of, 134-135, 138-139, Cretaceous-Paleocene, 74 reflection profiles
192 early Paleozoic, 68 Reservoirs
Cretaceous tectonics of, 188 of Eastern Venezuelan basin, 749-755 of Andes, 125-126
cross sections of, 153-158, 197, 200 Jurassic, 70-71 of Bolivia, 225
geologic maps of, 130-155, 187 late Jurassic–Neocomian, 72 of Cuyo basin, 366
hydrocarbon potential of, 198 late Paleozoic, 69 of Ecuador, 585-590
Phanerozoic tectonics of, 185-201 of Peru, 423-444 in foreland basins, 176-177
sedimentation of, 185-201 of sub-Andean belt, 423-444, 433-440 of Hollin Formation, 585-590
stratigraphic column of, 188, 198 Triassic, 69-70 Napo T and U sandstones, 581-582
tectonosedimentary framework of, Petroleum systems of Oriente basin, 585-590
190 of Bolivia, 451-457 of San Bernardo belt, 414
790 Index

sandstone, 125-126 regional map of, 404 of Marañon basin, 426


of sub-Andean belt, 433-436, 547-548 reservoirs of, 414 of Misoá Formation, 716
Resources, basin, 57-60 seismic reflection profiles of, 407, 408, of Nama Group, 12, 14
Retama Formation, as source rocks, 449 411 of Neuquén basin, 388, 389, 393, 395,
Retama Group, 216 stratigraphy of, 405-406 396
Rift basin structural inversion of, 403-413 of northeast-oriented lineaments, 22-
architecture of, 320 structure of, 405, 406-409, 413 23
of Eastern Cordillera, 639-639 San Jorge basin of Oriente basin, 563, 564, 565, 566,
of Maracaibo basin, 705-706 seismic reflection profile of, 411 567, 568, 569, 636, 637
seismic reflection profile of, 316-319 stratigraphic column of, 406 of Paganzo basin, 289
tectonics of, 639-639 San José Formation, stratigraphy of, 195 of rift basin, 316-319
Triassic, 69-70, 322 San Juan X2 well, maturity in, 457 of Salta rift, 327, 329, 330-331, 334-336
Rift-drift subsidence, definition of, 81 San Lorenzo Formation, stratigraphy of, of San Bernardo belt, 407, 408, 411
Rifts, types of, 80-81 619 of San Jorge basin, 411
Rio Apa subcraton, stratigraphy of, 192 San Luis, Neogene structure of, 351-352 of sub-Andean belt, 434-435, 549
Rio Chota basin, fluvial sequences of, San Marcos basin, stratigraphic column of Venezuelan Coastal Ranges, 671
603 of, 602 of western Venezuela, 685, 688, 692
Rio Grande area, cross section of, 379 San Mateo Formation, stratigraphy of, Sequence stratigraphy, see also Strati-
Rio Negro Formation, stratigraphy of, 623, 625 graphic columns; Supersequences
707 San Nicolás batholiths, 244 of Altiplano, 312
Riphean, extension in, 11-12 San Telmo Formation, stratigraphy of, of Boomerang Hills, 483-486
Rotational arc 259 of Eastern Venezuelan basin, 743-745
in Andean bend, 519-520 Sandstone reservoirs, in Andes, 125-126 of intermontane basins, 602-604
definition of, 512 Santa Barbara Formation, stratigraphy of Maracaibo basin, 702-704
Rumichaca basin of, 196-197 Sequences, definition of, 211
basin evolution of, 607-608 Santa Elena Formation, stratigraphy of, Serranía
cross section of, 610 619 cross section of, 672, 673
geologic map of, 610 Santa Elena Peninsula geologic map of, 670
stratigraphic column of, 603 paleocurrents of, 621, 624 Serranías Occidentales, see Occidental
setting of, 616-617 Sidewall, definition of, 20
Salado River area, cross section of, 377 stratigraphic column of, 620, 624 Sierra Chicas
Salamanqueano beds, 406 Santa Rosa Formation, stratigraphy of, Cretaceous stratigraphy of, 344-348
Salar de Atacama, correlation chart of, 190, 192 geologic map of, 345
321 Seca Formation, stratigraphic column of, Neogene structure of, 344
Saldán Formation 624 paleocurrents of, 347
geologic map of, 348 Seismic reflection profiles paleogeography of, 344-348
stratigraphy of, 347-348 of Aguada Villanueva, 395 structural inversion of, 348-351
Salta Group of Aguaragüe range, 534, 537 Sierra de Los Condores
angular unconformity of, 260 of Altiplano, 316-319 geologic map of, 347
paleogeography of, 263 of Andean structure, 27, 28 stratigraphy of, 346-347
tectonic evolution of, 276-277 of Bajada Vidal trough, 388 Sierra de Pajarillo
Salta rift of basin development, 25-28 cross section of, 346
explosive source data for, 339 of Boomerang Hills, 485, 486, 491, geology of, 345, 348
isopach map of, 326 492, 493, 495, 496 stratigraphy of, 345-346
postrift deposits of, 332-333 of Caipipendi, 540, 541 Sierra de Palauco area, cross section of,
prerift sequences of, 328 of Chaco basin, 15, 191 380
seismic reflection profiles of, 327, 329, of Charagüe, 537, 538 Sierra del Gigante
330-331, 334-336 of Cusiana field, 637 inversion of, 354-355
stratigraphy of, 328-333 of Cuyo basin, 364 slip analysis of, 353, 356
synrift deposits of, 330-332 of Eastern Venezuelan basin, 744, Sierra del Pajarillo, block diagram of, 346
tectonics of, 332, 326-328, 333-336 747, 748, 749, 750 Sierra Guayaguas, inversion of, 353-354
vibroseis data for, 339 of Entre Lomas trough, 389 Sierra Pampeanas, see Pampean Ranges
San Alfredo Group, stratigraphy of, 194- of Estancia Vieja area, 396 Sierra Quijadas, inversion of, 354
195 of Faja Petrolifera, 748 Sierras Chiquitanas, data point analysis
San Bernardo belt, 404 of Ghanzi thrusts, 12-13 of, 80
cross sections of, 407, 408, 409, 416 of Huincul dorsal, 38, 393, 395, 396 Sierras de Cordoba, geologic map of, 343
geodynamic interpretation of, 409- of hydrocarbon traps, 434-435 Single data point analysis, 80
412 of Izozog arch, 27 Slip sheet, gravity driven, 648-649
hydrocarbons of, 413-416 of Lomas de Olmedo rift, 264 Slow flexure, definition of, 81
oil fields of, 405 of Main sub-Andean thrust, 472 Snip restoration, of Eastern Cordillera,
oil-bearing anticline of, 415 of Maracaibo basin, 701, 715-717 476
Index 791

Socorro Formation, stratigraphic column of Eastern Venezuelan basin, 730, 745 cross sections of, 425, 447, 463, 533,
of, 624 of Ecuador, 601 541, 549
Socuy Member of Falcón basin, 760 crustal load deflection in, 463
isopach map of, 708 of glacial deposits, 168 crustal shortening in, 539-540
stratigraphy of, 707 of Guajira Peninsula, 760 deformation of, 535-539
Source rocks, see also Hydrocarbons of Guandacol supersequence, 290 evolution of, 74, 397-398
of Andes, 124-125 of Guayaquil area, 618 exploration modeling of, 450-451
of Bolivia, 448-450 of Hollin Formation, 576 foredeep fill of, 462
Copacabana Formation as, 449, 450 of Las Pavas supersequence, 256 gas-oil ratio in, 456
of Cuyo basin, 365-366 of Lomas de Olmedo basin, 40 geochemical evaluation of, 437, 526-
of Eastern Venezuelan basin, 730, 752 of Los Monos Formation, 529 531
Guayuta Group as, 737 of Maracaibo basin, 703, 704 geologic maps of, 464, 469, 525
Los Monos Formation as, 528-529, of Misoa Formation, 693 geologic setting of, 252-253, 446-448,
547 of Namora basin, 602 512, 524, 546-547
Retama Formation as, 449 of Neuquén basin, 372, 385 geothermal gradient of, 437
of San Bernardo belt, 413-414 of Occidental, 354 hydrocarbons of, 424, 433-440, 451-
of sub-Andean belt, 436, 448-450, 526- of Oriente basin, 562, 575, 576 457, 526-531, 547-549
530, 547 of Paganzo basin, 287 imbrication in, 464-466
Toregua Formation as, 450 of Pampean Ranges, 344 location maps of, 512, 532, 546
in upper Patquía, 293 of Peru, 602 main thrust tectonics of, 469-473
South America of Phanerozoic, 21, 253 oil and gas windows of, 451
architectural map of, 65 of Punta Ancón Formation, 624 oil–source correlation, 438
basin evolution in, 5-44 of Rumichaco basin, 603 petroleum geology of, 423-444, 433-
basin map of, 84, 167, 186, 286, 758 of San Jorge basin, 406 440, 452-455
and Caribbean plate boundary, 676 of San Marcos basin, 602 reservoirs of, 433-436, 547-548
geologic cross sections of, 153-156, of San Mateo Formation, 625 sedimentary wedge of, 449
157-158 of Santa Elena Peninsula, 620, 624 seismic reflection profiles of, 434-435,
geologic maps of, 130-155 of Seca Formation, 624 549
geologic provinces of, 131 of Socorro Formation, 624 source rocks of, 436, 448-450, 526-530,
geologic setting of, 64-66 of South America, 132-139 547
glacial record of, 167-168 of sub-Andean belt, 448, 462, 526, 548 stratigraphic columns of, 448, 462,
location map of northwestern, 634, of Tinajani basin, 603 526, 548
701 of Tupe supersequence, 291 stratigraphy of, 216, 426-430, 462-463,
oil and gas in, 53-61 of Vilcabamba basin, 601 524-526
paleogeography of, 67, 69 of western Venezuela, 684, 691, 694, structural trap styles of, 439-440
petroleum basins in, 54, 63-74 696 structure of, 447, 460-462, 463-464,
Phanerozoic correlation in, 129-158 Strike-slip fault, horizontal offset from, 531-541, 549-554
political map of, 460 653 synclines of, 467-468
tectonic features of, 701 Structural inversion tectonic framework of, 430-433
tectonic map of, 758 of Altiplano, 321 triangle zone of, 466-467
South American plate, margin of, 722, in Argentina, 341-356, 359-366 Triassic basin of, 429
757-778 cross section of, 321, 364 Subduction
South Caribbean deformed belt, 758 in Cuyo basin, 359-366 in Andes, 643
Southern hemisphere, paleogeography definition of, 35 in Eastern Venezuelan ranges, 675-
of, 86-96 in Huincul dorsal, 391-396 676
Stasis, definition of, 81 Jurassic-Cretaceous, 39 Subsidence curves
Strain partitioning, in Venezuelan Coast in Malargüe belt, 369-380 analysis of, 80-81
Ranges, 676 Neogene, 348-351, 353-356 for Carboniferous-Permian, 33
Stratigraphic columns in Neuquén basin, 369-380, 391-393 for La Valenciana half-graben, 373
of Altiplano, 309, 310, 311 in Pampean Ranges, 341-356 of Maracaibo basin, 709, 710, 715, 719
of Ayacucho basin, 602 of San Bernardo belt, 403-413 for NW Argentina, 262
of Bolivia, 213, 215, 219, 221, 448 of Sierra Chicas, 348-351 for Paganzo basin, 33, 295
of Boomerang Hills, 483 Sub-Andean belt for Venezuela, 776
of Chaco basin, 188, 194 and Altiplano, 474 Successor basins
of Chaco-Tarija basin, 178 Argentine ranges of, 547 definition of, 19
of Chongón-Colonche Cordillera, basin evolution of, 73-74 of Paleozoic, 19-34
620, 622 basin setting of, 425-426 Superpatagoniano succession, 406
of Chota basin, 601 burial history diagram of, 528 Supersequences
of coastal Ecuador, 618 and Chaco basin, 535-536 of Argentina, 253-261
of Cuenca basin, 601 correlation chart of, 427, 431 of Bolivia, 209
of Cuyo basin, 40, 361 Cretaceous of, 427, 430 definition of, 211
792 Index

of Paganzo basin, 288-296 definition of, 648, 650 crust of, 667-677
of western Venezuela, 683, 685-692 of sub-Andean belt, 466-467 exploration efficiency for, 58
Trujillo Formation geologic history of northwestern, 771-
Tacurú supersequence, stratigraphy of, clinoforms of, 715 774
259-260 progradation in, 717 isopach maps of, 772, 773, 775
Tarija basin, stratigraphy of, 259 seismic reflection profile of, 715 location maps of, 648, 669
Tarija Formation stratigraphy of, 712 seismicity of, 669
outcrops of, 179 Tuichi syncline, cross section of, 473 stratigraphy of, 681-697, 759-774
stratigraphy of, 258 Tupambi Formation, stratigraphy of, 258 tectonics of, 754, 758-759, 669
Tarija-Teoponte belt, stratigraphy of, Tupe supersequence, stratigraphy of, wildcat drilling in, 54-55
216-218 290-291, 292, 293 Venezuelan Coast Ranges, see also
Tecka Range, stratigraphic column of, Turbidites Eastern Venezuelan ranges
171 depositional model of, 176 basement shortening in, 674
Tectonic subsidence, see also Subsidence of Tarija Formation, 179 cross section of, 672, 673
curves of Tupe supersequence, 293 crust of, 667-677
curves of, 80-81, 262, 295, 373 Two-dimensional computer modeling, geologic setting of, 668-672
of Cuyo basin, 363 727-740 geophysical modeling of, 675
of Gondwana, 79-96 gravimetric map of, 673
of Maracaibo basin, 715, 719 Ucayali basin magnetic map of, 674
Tectonic wedge correlation chart of, 427 seismic reflection profile of, 671
creating trap by, 552 exploration wells of, 442 strain partitioning in, 676
delamination by, 551 hydrocarbons of, 424, 433-440 Vibroseis data, for Salta rift, 339
high-density in Venezuelan Coast oil fields of, 443 Vilcabamba basin, stratigraphic column
Ranges, 676-677 setting of, 426 of, 601
TEMISPACK software, 728, 736, 737 Uncía Formation, stratigraphy of, 213 Vilque well, stratigraphic column of, 309
Temperature-burial modeling, for San Undiscovered resources, prediction of, Vitiacua Formation, stratigraphy of, 259
Bernardo belt, 415 60 Vivian Formation, isopach map of, 432
Tepuel basin Upper Hollin Formation, see also Hollin Vizcacheras block, structure of, 365
glacial deposits of, 170-171 Formation
stratigraphic framework of, 169 depositional systems of, 577-579 Wells, wildcat, 54-55
Tepuel Group, stratigraphic column of, Mohr-Coulomb failure criteria of, 596 Western Cordillera
171 petrophysical properties of, 591, 592 geologic setting of, 512-514
Tequeje Formation, stratigraphy of, 502- porosity-permeability of, 590 subduction on, 643
503 Upper Patquía–De la Cuesta superse- Western Venezuela
THEMIS software, 751 quence, 292, 293 correlation chart of, 684, 703
Tien Shan mountains, gravity-driven Uribante rift, Jurassic of, 706 geologic setting of, 682
structures in, 649 Urica fault, seismic reflection profile of, logs of, 694, 696
Tinajani basin, stratigraphic column of, 748 paleogeography of, 683, 688-692, 694-
603 Uruguay 696
Tobifera, 405 correlation chart of, 136-137 seismic reflection profiles of, 685, 688,
Toe addition, definition of, 465 geologic cross sections of, 153-158 692
Tokochi Formation, stratigraphy of, 213 geologic maps of, 130-155 stratigraphic column of, 684, 691, 694,
Tomachi Formation, stratigraphy of, 503 Urumaco trough 696
Tordillo Formation, paleogeography of, interpreted seismic line of, 763 stratigraphy of, 682-696
392 subsidence rates of, 776-777 tectonic evolution of, 683, 700-723
Toregua Formation tectonic setting of, 700-702
as source rock, 450 Vaca Muerta range, outcrops of, 394 Wildcat drilling
stratigraphy of, 503 Venezuela, see also Western Venezuela in Latin America, 54-55
Transpression, Permian-Triassic, 36 Caribbean pull-apart basin of, 757- net oil found by, 57
Traps, see Hydrocarbons 778
Triangle zone Cenozoic tectonics of, 757-778 Yaurichambi Formation, stratigraphy of,
cross section of, 649 cross section of, 754 503

You might also like